UC San Diego UC San Diego Electronic Theses and Dissertations

Title Heterogeneous and multiphase chemistry of reactive nitrogen oxides in the marine boundary layer

Permalink https://escholarship.org/uc/item/5zd857fz

Author Ryder, Olivia Sharmelle

Publication Date 2015

Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital Library University of California UNIVERSITY OF CALIFORNIA, SAN DIEGO

Heterogeneous and multiphase chemistry of reactive nitrogen oxides in the marine boundary layer

A dissertation submitted in partial satisfaction of the requirements for the degree Doctor of Philosophy

in

Chemistry

by

Olivia Sharmelle Ryder

Committee in charge:

Professor Timothy Bertram, Chair Professor Lihini Aluwihare Professor Meinrat Andreae Professor Patricia Jennings Professor Francesco Paesani Professor Michael Tauber

2015

Copyright

Olivia Sharmelle Ryder, 2015

All rights reserved.

The Dissertation of Olivia Sharmelle Ryder is approved, and is acceptable in quality and form for publication on microfilm and electronically:

______

______

______

______

______

______

Chair

University of California, San Diego

2015

iii

DEDICATION

To my parents- thank you for your love, sacrifices,

and for always believing in me.

iv

EPIGRAPH

If research was easy, everyone would be doing it -B. J. Finlayson-Pitts

v

TABLE OF CONTENTS

Signature Page...... iii

Dedication...... iv

Epigraph...... v

Table of Contents...... vi

List of Figures...... ix

List of Tables...... xii

Acknowledgements...... xiii

Vita...... xvi

Abstract of The Dissertation...... xviii

Chapter 1 Introduction ...... 1

1.1 The Role of N2O5 in the Atmosphere ...... 2

1.2 N2O5 Hydrolysis ...... 3

1.3 The Reactive Uptake Probability, (N2O5) ...... 4

1.4 Particle Chemical Composition and N2O5 Reactive Uptake ...... 5 1.4.1 The Effect of Chemical Composition of Ambient Particles: Field Measurements of (N2O5) ...... 5 1.4.2 Uptake to Inorganic Salts: Laboratory Studies ...... 6 1.4.3 The Impact of Particulate Organics: Laboratory Studies ...... 6 1.4.4 Laboratory and Field Measurement Agreement ...... 7

1.5 Product Formation Following N2O5 Reaction on Chloride-Containing Surfaces ...... 8

1.5.1 Ambient Measurements of ClNO2 and (ClNO2) ...... 9 1.6 Summary and Remaining Areas to Address ...... 10 1.7 Synopsis of Chapters ...... 12 1.8 References ...... 13

Chapter 2 On The Role of Particle Inorganic Mixing State in The Reactive Uptake of N2O5 to Ambient Aerosol Particles ...... 17 2.1 Abstract ...... 17 2.2 Introduction ...... 17

2.2.1 Factors Controlling the Reactive Uptake of N2O5 to Aerosol Particles ...... 18

vi

2.2.2 Single Particle Chloride-to-Nitrate Mixing State in The Polluted Marine Boundary Layer ...... 19 2.3 Materials and Methods ...... 21 2.4 Results and Discussion ...... 22

2.4.1 Impact of Particulate Chloride-to-Nitrate Mixing State on (N2O5) ...... 22 2.4.2 Observation-Based Constraints on the Impact of Chloride-to-Nitrate Particle Mixing State on (N2O5) ...... 26 2.5 Insight from Single Particle Measurements of Reacted Sea Spray Aerosol ...... 31 2.6 Atmospheric Implications ...... 34 2.7 Acknowledgements...... 35 2.8 Supporting Information ...... 35 2.8.1 Sampling Locations and Details ...... 35 2.8.2 Measurements ...... 37 2.8.3 Calculations and Interpretations ...... 40 2.9 Acknowledgments ...... 48 2.10 References ...... 48

Chapter 3 On the Role of Organic Surfactants and Coatings in Regulating N2O5 Reactive Uptake to Sea Spray Aerosol ...... 53 3.1 Abstract ...... 53 3.2 Introduction ...... 54 3.3 Materials and Methods ...... 57

3.3.1 N2O5 Generation and Detection ...... 57 3.3.2 Sea Spray Aerosol Generation and Duty Cycle ...... 58 3.3.3 Molecular Mimics for a Synthetic Phytoplankton Bloom ...... 60 3.4 Wave Channel Generated Sea Spray ...... 63 3.5 Seawater and Sea Spray Aerosol Characterization ...... 64 3.5.1 Aerosol Particle Characterization ...... 64 3.5.2 Water Characterization ...... 66

3.6 (N2O5) Determinations ...... 66 3.7 Results and Discussion ...... 70 3.7.1 Aerosol Organic Volume Fraction ...... 70

3.7.2 Organic Addition ESffect on (N2O5) ...... 72 3.8 Conclusions ...... 76 3.9 Acknowledgements...... 77 3.10 References ...... 77

Chapter 4 On the Role of Organics in Regulating ClNO2 Production at the Air-Sea Interface 83

vii

4.1 Abstract ...... 83 4.2 Introduction ...... 83 4.3 Materials and Methods ...... 86

4.3.1 N2O5 Generation and Sample Exposure ...... 86 4.3.2 Ambient Sample Collection ...... 87 4.3.3 Laboratory Organic Mimic Preparation ...... 88 4.3.4 Surface Tension Measurements ...... 89 4.3.5 Box Model ...... 89 4.4 Results and Discussion ...... 90 - 4.4.1 (ClNO2) Dependence on Cl ...... 90

4.4.2 ClNO2 Yields from Ambient Seawater Samples ...... 92

4.4.3 (ClNO2) from Laboratory Mimics ...... 95 4.5 Linking Laboratory and Field Studies...... 101 4.6 Summary and Conclusions ...... 103 4.7 Acknowledgements...... 104 4.8 References ...... 105 Chapter 5 The potential role of divalent cations at the air-sea interface in regulating the production rate of ClNO2 ...... 109 5.1 Abstract ...... 109 5.2 Introduction ...... 109 5.3 Materials and Methods ...... 111 5.3.1 Solution Preparation ...... 111 5.3.2 Real and Proxy Ocean Water Solutions ...... 111

5.3.3 N2O5 and ClNO2 detection ...... 112

5.3.4 N2O5 generation and delivery ...... 112 5.3.5 Ion Chromatography Analysis ...... 113 5.3.6 Solution pH and Surface Tension Measurements ...... 113 5.3.7 Inductively Coupled Plasma-Mass Spectrometry (ICP-MS) Analysis ...... 114 5.4 Results and Discussion ...... 114 5.5 Future work/ challenges ...... 124 5.6 Acknowledgments ...... 124 5.7 References ...... 125

viii

LIST OF FIGURES

Figure 1.1 The overlap between the NOx and O3 cycles in the troposphere...... 2

Figure 1.2 Box model results showing the fraction of NOx lost at night due to N2O5 and NO3 reaction...... 3

- Figure 1.3 ClNO2 yield following N2O5 reaction as a function of [Cl ] in solution from which particles were atomized. The results shown are from 3 separate studies...... 9

Figure 1.4 Model calculations illustrating the dependence of NOx loss (left y axis) and chlorine radical production following photolysis of ClNO2 (right y axis) on the magnitude of (N2O5). This plot has been adapted from Bertram et al. 2009.20 ...... 11

Figure 2.1 A) Model parameterization of (N2O5), where nascent SSA are emitted at high - - - nH2O/nNO3 (aq) and nCl /nNO3 , and exhibit large (N2O5) in the absence of organic films. As - - nascent SSA is chemically processed in the atmosphere, nCl /nNO3 decreases and the rate of N2O5 uptake slows...... 24

Figure 2.2 A) Backward air trajectories calculated for 24 hours prior to sampling for Period 1 and Period 2. B) Particle number concentration (N, left axis; see supporting information for sizing measurement details) and NOx (right axis)...... 25

Figure 2.3 A) Histogram of calculated single particle (N2O5), based on ATOFMS single particle - - measurements of nCl /nNO3 at the SIO pier, for sampling period 1 (more-polluted; pink) and period 2 (less-polluted; blue)...... 29

Figure 2.4 Probability distribution functions (PDF) of (A) fROA (calculated using the representative organic ions, m/Q (37 + 43) / total ion peak area) and (B) the contribution of m/Q 43 to the organic signal (m/Q 43 / m/Q (37 + 43) ) for the more-polluted (red) and less-polluted (blue) pier data sets...... 30

- - Figure 2.5 Probability distribution function (PDF) as a function of nCl /nNO3 for: A) more- polluted pier data, B) less-polluted pier data, and C) nascent SSA data sets. D) Single particle - - nCl /nNO3 as a function of particulate organic fraction (fROA) for nascent SSA...... 32

- - Figure 2.6 Probability distribution curves of particulate nCl /nNO3 calculated for four cases: ambient aerosol from Jeju Island, Korea (grey) and Long Beach, CA (black), nascent sea spray aerosol generated in the SIO wave channel (blue), and the same nascent sea spray after reaction with HNO3 (red)...... 42

Figure 2.7 Assessment of variation in the cumulative distribution function arising from sources - - unrelated to particle-to-particle variability. Panel A shows the raw variability in nCl /nNO3 over 24 hours of sampling at Jeju Island (light grey markers). The running median over 500s of sampling is shown in blue ...... 45

Figure 3.1 A typical surface area distribution for SSA produced in the MART system after applying a growth factor value of 2.3, and adjusting for dilution of the SSA concentration to the sizing instrumentation so as not to saturate the detectors...... 60

ix

Figure 3.2 Plunging duty cycle flag as a function of time (Panel A) where a value of 1 indicates plunging is turned on for SSA production mode, and a value of zero indicates plunging is turned off. Panel B shows the response of the total (red) and sub-micrometer (green) surface area to changes in the plunging duty cycle as a function of time...... 61

Figure 3.4 Panel A shows a representative time trace of total particle surface area (grey) and super-micrometer surface area (green) during a typical experimental cycle, where time=0 corresponds to plunging turned off, thus switching from SSA production mode to SSA decay mode...... 68

Figure 3.5 Representative plot of instantaneous kobs versus the total surface area in the flow reactor. Data was filtered such that over the length of one period of residence time in the flow reactor, the surface area was not changing by more than 20 % (see text for details)...... 70

Figure 3.6 AFM derived organic volume fraction of 0.33-0.56 m particles (measured as 0.21 m at 21 % RH by AFM) collected using a MOUDI, imaged for each water matrix vs. the fraction of occurrence...... 71

Figure 3.7 (N2O5) as a function of water side analysis metrics including carbon concentration added to the water matrix (A), water pH (B), surface tension (C). All points are colored by the amount of carbon present in the water matrix, and the real ocean sample is colored green...... 73

Figure 3.8 (N2O5) as a function of the organic mass fraction from this work (green, color coded by TOC water content) as compared to values obtained from a real mesocosm bloom generated in the wave channel facility at Scripps Institution of in 2011 (yellow) ...... 76

Figure 4.1 Signal intensities for N2O5 (top panel) and ClNO2 (bottom panel) during N2O5 delivery to CI-QMS inlet (I), after flowing over a saturated NaCl solution (6.1 M NaCl) (II), and back to the CI-QMS inlet (III)...... 91

Figure 3.2 ClNO2 yield as a function of chloride concentration as determined in this work (red squares) and fit values from previous determinations, Roberts et al., 2009 (circles), Bertram and Thornton, 2009 (inverted triangles), and Behnke et al., 1997 (diamonds)...... 92

Figure 4.3 Dependence of the measured ClNO2 yield on chlorophyll-a, a proxy for biological activity, for ambient sea-water samples collected from coastal and open ocean water and from a large scale mesocosm bloom experiment...... 96

Figure 4.4 Surface tension of laboratory samples as a function of [phenol]bulk (A, red squares) along with a linear fit to the data (dashed line). Calculated phenol surface excess (B, blue circles), as derived from the linear fit shown in Panel A, and the calculated concentration of chloride at the solution interface (dashed line)...... 98

Figure 4.5 ClNO2 yield determinations (red squares) as a function of the molar ratio of phenol to chloride at the surface (bottom axis) and in the bulk (top axis) following sequential additions of phenol to 500 mM NaCl solution...... 101

x

Figure 4.6 ClNO2 yield as a function of carbon concentration, assuming homogenous mixing, following sequential additions of cholesterol (blue squares) and humic acid (black circles) into a 500 mM NaCl solution...... 103

Figure 5.2 ClNO2 yield following MgCl2 anhydrous addition to a 500 mM NaCl solution (circles, panel A). The solid line indicates the typical concentration of Mg2+ in the ocean. Panel B shows 2+ the ClNO2 yield following EDTA addition to a solution of 53 mM Mg in 500 mM NaCl solution...... 116

Figure 5.3 ClNO2 yield as a function of EDTA molar equivalent additions to 53 mM MgCl2 anhydrous (blue) and artificial reef salt (green). The dashed lines are fits to serve as a guide. ... 117

2+ Figure 5.4 ClNO2 yield as a function of [Mg ] added to a 500 mM NaCl solution for both MgCl2.6H2O (grey squares) and MgCl2 anhydrous (red squares). The oceanic concentration of magnesium is denoted by the blue line at 53 mM Mg2+...... 119

Figure 5.5 ICP-MS analysis of coastal ocean water (A), and artificial reef salt (B), MgCl2.6H2O, Sigma Aldrich (C), MgCl2.6H2O, Macron (D), MgCl2 anhydrous (E)...... 122

Figure 5.6 A closer view of Figure 5.5, showing ions present in lower concentrations. Inset Panel F shows a closer view of the 57Fe peak signal...... 123

xi

LIST OF TABLES

Table 2.1N2O5 values calculated using the assumption of an externally mixed particle population, an internally mixed population, and measured from SIO pier, along average RH and particle properties for sampling Periods 1 and 2...... 27

Table 2.2 The fraction of particles removed before CDF and/or N2O5 calculation...... 40

Table 3.1 Synthetic seawater components and their respective properties...... 63

Table 3.2 Average gamma values obtained during each stage of the synthetic bloom in addition to the range of relative humidities inside the flow reactor over which gamma was measured, and the pH and surface tension of the water matrix...... 66

Table 4.1 Bulk carbon concentrations following sequential addition of organic molecular mimics to 500 mM NaCl solutions...... 89

Table 4.2 Ambient water samples, corresponding ClNO2 yields, and co-located measurements of chlorophyll-a, colored dissolved organic matter (CDOM), dissolved organic carbon (DOC), and surface tension where available...... 94

Table 5.1 Sample solutions and their associated measured pH, [Mg2+] as measured by IC, and surface tension...... 118

xii

ACKNOWLEDGEMENTS

I would like to thank Timothy Bertram, my thesis advisor, for taking me on as his first graduate student and for his training, encouragement, support and for affording me many invaluable research opportunities.

I would also like to thank all members of the Bertram group, from when we were at our strongest (Timia, James, Michelle, Nicole, Steve, Katy, Matt), for great times spent together in the lab/office and during field studies, for your support during hard times, and for constant comic relief! There was never a dull moment, and you all helped make my graduate school experience a great one- thank you so much!

Additionally, Kim Prather and everyone involved in CAICE since its beginning have had a profound impact on my development as a researcher and in my ability to collaborate, and for this I would also like to thank you all.

Last but not least, my family and friends who have stuck by me all throughout my education, and especially over the last 6 years - thank you so much. Without you all backing me, cheering me on, and supporting me the whole way through good times and bad, I would not have been able to achieve this.

Chapter 2, in full, is a reprint of the material as it appears in Environmental Science and

Technology. Olivia S. Ryder, Andrew P. Ault, John F. Cahill, Timothy L. Guasco, Theran P.

Riedel, Luis A. Cuadra-Rodriguez, Cassandra J. Gaston, Elizabeth Fitzgerald, Christopher Lee,

Kimberly A. Prather, and Timothy H. Bertram. (2014), On the Role of Particle Inorganic Mixing

State in the Reactive Uptake of N2O5 to Ambient Aerosol Particles, Environ. Sci. Technol., 2014,

48 (3), pp 1618–1627, doi: 10.1021/es4042622. The dissertation author was the primary investigator and author of this paper.

xiii

The authors would like to thank Joel Thornton for the use of the CIMS and flow tube apparatus used in the ambient portion of this study. This research was supported by the Office of

Science (Office of Biological and Environmental Research), U.S. Department of Energy (Grant

No. DE-SC0006431) and the National Science Foundation (Grant No. CHE1038028). O.S.R is grateful for a Graduate Research Fellowship from the National Science Foundation (2011-2014).

Chapter 3, in part, is currently being prepared for submission of the material.

Olivia S. Ryder, Nicole R. Campbell, Holly Morris, Matthew J. Ruppel, Alexei Tivanski, and

Timothy H. Bertram. The dissertation author was the primary investigator and author of this material.

This research was supported by the National Science Foundation via the Center for

Aerosol Impact on Climate and the Environment, a Center for Chemical Innovation (NSF

CHE1305427). The authors gratefully acknowledge Dr. Chris Cappa for use of SMPS and growth factor data used in this study and for helpful discussions throughout the experiments. The authors also thank Steven Schill and Sara Forestieri for participating in the collaborative effort of running these experiments. O.S.R gratefully acknowledges a Graduate Research Fellowship from the

National Science Foundation (2011-2014).

Chapter 4, in full, is currently under review in The Journal of Physical Chemistry A.

O.S. Ryder, N.R. Campbell, M. Shaloski, H. Al-Mashat, G. M. Nathanson, and T.H. Bertram.

(2014), On the Role of Organics in Regulating ClNO2 Production at the Air-Sea Interface. The dissertation author was the primary investigator and author of this material.

This research was supported by the National Science Foundation via the Center for

Aerosol Impact on Climate and the Environment, a Center for Chemical Innovation (NSF

CHE1305427). The authors gratefully acknowledge Dr. Grant Deane and Dale Stokes for use of the tensiometer used in this work. Dr. Matthew Zoerb (UCSD) for the collection of Western

Atlantic Cruise II 2014 water samples, Bob Vaillancourt, Jeremiah Stone, and Evan Ntonados,

xiv

(all from Millersville University of Pennsylvania) for providing WACS2 chlorophyll-a concentrations, and the entire CAICE IMPACTS team and the staff of the SIO Hydraulics

Laboratory. O.S.R gratefully acknowledges a Graduate Research Fellowship from the National

Science Foundation (2011-2014).

Chapter 5, in part, is currently being prepared for submission of the material.

Olivia S. Ryder and Timothy H. Bertram. The dissertation author was the primary investigator and author of this material.

This work was supported by the National Science Foundation via the Center for Aerosol

Impact on Climate and the Environment, a Center for Chemical Innovation (NSF CHE1305427).

The authors would like to gratefully thank Scott Wilson and Dr. Skip Pomeroy for IC analysis,

Dr. James Day for ICP-MS analysis, Dr. Dale Stokes and Dr. Grant Deane for use of the surface tensiometer used in this work, Dr. Kim Prather, Dr. Michael Tauber and Dr. Andreae Meinrat for helpful discussions. O.S.R gratefully acknowledges a Graduate Research Fellowship from the

National Science Foundation (2011-2014).

xv

VITA

2006-2008 Undergraduate Research Assistant, Department of Chemistry University of California, Irvine

2008 Bachelor of Science, Chemistry (Honors), Magna Cum Laude, University of California, Irvine

2009-2010 Teaching Assistant, University of California, San Diego

2009-2015 Research Assistant, Department of Chemistry University of California, San Diego

2011-2014 National Science Foundation Graduate Research Fellowship Program Fellow

2011 Master of Science, University of California, San Diego

2015 Doctor of Philosophy, University of California, San Diego

PUBLICATIONS

Ryder, O. S.; Ault, A.; Cahill, J. F.; Guasco, T. L.; Cuadra-Rodriguez, L.; Thornton, J. A.; Gaston, C.; Fitzgerald, E.; Lee, C.; Prather, K. A.; Bertram, T. H. On the Role of Particle Mixing State in the Reactive Uptake of N2O5 to Ambient Aerosol Particles. Environ. Sci. Technol., 2014, 48(3), 1618-27

Ault, A. P.; Guasco, T.L. ; Baltrusaitis, J.; Ryder, O. S.; Trueblood, J.; Collins, D. B.; Ruppel, M. J.; Cuadra-Rodriguez, L. A.; Prather, K. A.; Grassian, V. H. Heterogeneous Reactivity of Nitric Acid with Nascent Sea Spray Aerosol: Large Differences Observed Between and Within Individual Particles. J. Phys. Chem. Lett., 2014, 5, 2493-2500

Ault, A. P.; Guasco, T. L.; Ryder, O. S.; Baltrusaitis, J.; Cuadra-Rodriguz, L. A.; Collins, D. B.; Ruppel, M. J.; Bertram, T. H.; Prather, K. A.; Grassian, V. H. Inside versus Outside: Ion Redistribution in Nitric Acid Reacted Sea Spray Aerosol Particles as Determined by Single Particle Analysis. J. Am. Chem. Soc. - Communication, 2013, 135(39), 14528-14531.

Ebben, C.; Ault, A.; Ruppel, M.; Ryder, O.; Bertram, T.; Grassian, V.; Prather, K.; Geiger, F. Size-Resoslved Sea Spray Aerosol Particles Studied by Vibrational Sum Frequency Generation. J. Phys. Chem. A, 2013, 117 (30), 6589–6601

Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; DeMott, P. J.; Aluwihare, L. I.; Palenik,B.; Azam, F.; Seinfeld, J. H.; Moffet, R. C.; Molina, M. J.; Cappa, C.

xvi

D.; Geiger, F. M.; Roberts, G. C.; Russell, L. M.; Ault, A. P.; Baltrusaitis, J.; Collins, D. B.; Corrigan, C. E.; Cuadra-Rodriguez, L. A.; Ebben, C. J.; Forestieri S. D.; Guasco, T. L.; Hersey, S. P.; Kim, M. J.; Lambert, W.; Modini, R. L.; Mui, W.; Pedler, B. E.; Ruppel, M. J.; Ryder, O. S.; Schoepp, N.; Sullivan, R. C.; Zhaoa, D. Bringing the Ocean into the Laboratory: Impacts of Chemical Complexity of Sea Spray Aerosol on Climate Properties. Proc. Nat. Acad. Sci. USA, 2013, 110(19), 7550-7555

Stokes, M. D.; Deane, G. B.; Prather, K.; Bertram, T. H.; Ruppel, M. J.; Ryder, O. S.; Brady, J. M.; and Zhao, D. A Marine Aerosol Reference Tank System as a Breaking Wave Analogue. Atmos. Meas. Tech., 2013, 6, 1085-1094

Riedel, T. P.; Bertram, T. H.; Ryder, O. S.; Liu, S.; Day, D. A.; Russell, L. M.; Gaston, C. J.; Prather, K. A.; and Thornton, J. A. Direct N2O5 Reactivity Measurements at a Polluted Coastal Site. Atmos. Chem. Phys., 2012, 12, 2959-2968

Bertram, T. H., Kimmel, J. R., Crisp, T. A., Ryder, O. S., Yatavelli, R. L. N., Thornton, J. A., Cubison, M. J., Gonin, M., and Worsnop, D. R. A Field-Deployable, Chemical Ionization Time- of-Flight Mass Spectrometer. Atmos. Meas. Tech., 2011, 4, 1471-1479

FIELDS OF STUDY

Major Field of Study: Chemistry (Atmospheric & Analytical Chemistry)

xvii

ABSTRACT OF THE DISSERTATION

Heterogeneous and multiphase chemistry of reactive nitrogen oxides in the marine boundary layer

by

Olivia Sharmelle Ryder

Doctor of Philosophy in Chemistry

University of California, San Diego, 2015

Professor Timothy Bertram, Chair

Dinitrogen pentoxide (N2O5), a primary nocturnal reservoir for NOx (NOx ≡ NO + NO2), readily undergoes heterogeneous reaction on aqueous interfaces. These reactions potentially alter the chemical, physical, and hygroscopic properties of ambient particles, and ultimately their climate impact. Upon reacting with aqueous chloride-containing surfaces, such as sea spray,

xviii

nitryl chloride (ClNO2) is produced. ClNO2 is capable of extending the lifetime of NOx and inducing ozone production in high NOx environments via chlorine radical creation. The extent to which N2O5 reactivity impacts atmospheric oxidant loadings depends on the heterogeneous reaction rate of N2O5 with a given surface. This, in turn, is directly influenced by interfacial chemical composition, morphology, and phase state.

In this dissertation, I describe advances in the understanding of how organic and inorganic constituents at atmospheric interfaces regulate N2O5 reactivity.

First, I developed an observation-based framework for assessing the aerosol particle mixing state role on N2O5 heterogeneous reaction kinetics using simultaneous measurements of the N2O5 reactive uptake coefficient to ambient aerosol with direct measurements of aerosol mixing state before and after reaction. Second, as part of the NSF Center for Aerosol Impacts on

Climate and the Environment, I extended the use of a novel Marine Aerosol Reference Tank

(MART) capable of generating nascent sea-spray aerosol, to determine heterogeneous reaction kinetics. Here, the MART was doped with ocean organic mimics to assess how organics present in the sea surface microlayer (SSML), and subsequently ejected into sea-spray, impact N2O5 reaction kinetics. Third, I conducted laboratory studies to investigate suppressed ClNO2 production rates at the ocean surface, using ambient seawater from various locations and under varied biological conditions, along with model sea salt solutions with proxy organics. This study was designed to determine the role of unsaturated organics and surface-enhanced inorganic ions in the SSML in altering ClNO2 production rates.

xix

Chapter 1 Introduction

The tropospheric atmosphere is comprised of two main classes of components: aerosol particles (or simply aerosol) and gases. Aerosol are defined as liquid or solid particles suspended within a gas.1 It is critical to understand aerosol and their chemical and physical changes throughout their lifetime in the atmosphere, as they have the ability to impact visibility, human health, air quality, and climate.1 More specifically, aerosol can impact climate by acting as cloud condensation nuclei (CCN), by which they serve as a surface upon which water vapor can condense to form a cloud droplet.2 However, the size and composition of aerosol particles will dictate their ability to act as CCN, and as such it is important to understand these complex components of the atmosphere. Aerosol vary in size and composition, both of which are initially linked to the production mechanism and source of the particle.1,2 However, throughout their lifetime, these particles undergo atmospheric processing, including adsorbing and desorbing water, coagulation, or reaction with gas-phase molecules, capable of altering their chemical composition. The latter process is referred to as heterogeneous chemistry and is the main focus of this dissertation.

Heterogeneous chemistry involves chemical reactions that take place between species of different phases, such as between a gas molecule and an aerosol particle. These reactions are important not only in terms of aerosol processing, potentially making particles more or less hygroscopic, but also due to the reactive gases these reactions can remove from the atmosphere.

Further, these reactions are capable of producing reaction products, which may partition into the gas phase, and undergo further reaction. Here we will focus on the heterogeneous reaction that takes place between dinitrogen pentoxide (N2O5) and reactive surfaces available in the marine environment.

1

2

1.1 The Role of N2O5 in the Atmosphere

N2O5 is an important tropospheric nitrogen-containing nocturnal species, that is formed from the reaction of NO2 and NO3. Its formation and reaction are limited to the nighttime hours as

1 it is subject to thermal decomposition, and photodissociates between =200 - 400 nm. NO3 is also photolabile, dissociating within the visible range (410-670 nm), resulting in a lifetime of approximately 6 seconds. The diminished availability of NO3 greatly limits the subsequent generation of N2O5 during sunlight hours.

N2O5 is considered a nocturnal terminal sink for NOx (NOx=NO + NO2) species in the atmosphere as upon reaction with an aqueous surface, it produces HNO3, thus removing NOx

3,4 from the atmosphere. This in turn has a direct impact on the ozone (O3) and hydroxyl radical

(OH) budgets as shown in Figure 1.1.

Figure 1.1 The overlap between the NOx and O3 cycles in the troposphere.

At night, NO2 will react with O3 to produce NO3, a key oxidant in the nocturnal atmosphere, that establishes equilibrium with N2O5. Therefore, the fate of N2O5 overnight will dictate the oxidative capacity of the atmosphere during this period. Additionally, if N2O5 persists

3

at dawn, it will thermally decompose back into NO2 and NO3. However, if it has reacted during the night, a lower concentration of NO2 will be present during the morning hours, effecting subsequent O3 and OH generation. It is therefore of key importance to understand N2O5 reactions to accurately predict the ozone, OH and NOx budgets in the polluted marine atmosphere. Figure

1.2 shows a model calculation illustrating that a significant fraction (here, 60 %) of NOx present at sundown can be removed via N2O5 reaction at night.

Figure 1.2 Box model results showing the fraction of NOx lost at night due to N2O5 and NO3 reaction.

1.2 N2O5 Hydrolysis

E.g.5–12 N2O5 hydrolysis has been studied on a variety of surfaces and the accepted mechanism is shown in Reactions 1.1-1.6.7,13,14

N2O5 (g) ↔ N2O5 (aq) (R1.1)

+ - N2O5 (aq) + H2O(l) ↔ H2ONO2 (aq) + NO3 (aq) (R1.2)

+ + H2ONO2 (aq) + H2O (l) ↔ H3O (aq) + HNO3 (aq) (R1.3)

+ - H2ONO2 (aq) + X ↔ XNO2 (aq) +H2O (l) (R1.4)

4

XNO2 (aq) ↔ XNO2 (g) (R1.5)

After N2O5(g) accommodation to the particle surface, autoionization allows for

+ - formation of nitronium ion (NO2 ) and nitrate (NO3 ), shown in Reactions 1.1-1.2. If the particle

+ - + contains a high concentration of nitrate (where k(NO2 + NO3 ) > k(NO2 + H2O)) the ionization

13,14 reaction will be driven in the reverse direction. For example, the probability of N2O5 reaction

- with aerosol decreases by a factor of 90 when [NO3 ] aerosol loading increases from 13.4 to 158 mol/kg.13–15 The fate of the nitronium ion is determined by the composition of the particle.19,20

+ NO2 will either react with water, forming nitric acid (Reaction 1.3) or, if the particle contains a halide ion (X-), will react to form a nitryl halide species.8,16–18 The nitryl halide species, depending on its Henry's Constant value (KH), can partition out of the aqueous phase into the gas phase. It is important to note that Reaction 1.3 represents a terminal loss of NOx in the nocturnal atmosphere.

1.3 The Reactive Uptake Probability, (N2O5)

When N2O5 collides with a surface, it does not necessarily react. The ability of N2O5 to undergo hydrolysis within a particle depends strongly on the composition of the particle, as will be described in the next section. The metric used to determine the extent of N2O5 reaction with a surface is given by the reactive uptake coefficient, (N2O5), which can also be thought of as the probability of an N2O5 molecule colliding with a surface and reacting (Equation 1.1):

(E1.1)

khet is the pseudo first order rate coefficient for N2O5 loss to particles,  is the mean molecular velocity of N2O5, and SA is the surface area of the particle population with which N2O5 is reacting. Within this expression,  accounts for a number of physical and chemical processes that

5 are involved in the "uptake" of a gas to a surface, including diffusion and gas-surface collision frequency, the mass accommodation involved in the gas crossing the interface, solubilization and diffusion into the bulk, and reaction within the particle.2 Laboratory measurements suggest that

N2O5 hydrolyzes within the first 50 nm of the air-particle interface, and thus is considered more a

6,7 surface rather than bulk reaction. As such N2O5 reactivity is strongly affected not only by the chemical composition of particles,19,20 but is acutely sensitive to the composition at the surface layer, where the reaction takes place. The chemical composition of particles influencing (N2O5) is a primary focus of this dissertation.

1.4 Particle Chemical Composition and N2O5 Reactive Uptake

1.4.1 The Impact of Chemical Composition: Field Measurements of (N2O5)

Ambient measurements of (N2O5) have been reported from a number of areas within the

United States and show a wide range of variability in measured (N2O5). Ambient reactive uptake studies in a polluted coastal location showed significant day to-day (N2O5) variability (<0.001-

0.029), which the authors attribute to varying levels of particulate nitrate content.21 Wagner et al. observed a similar reliance on nitrate loadings during wintertime studies in a continental region,

22 where they measured (N2O5) range between 0.02-0.1. Bertram et al. made direct measurements of (N2O5) on ambient aerosol particles, correlating the values to the particulate organic matter

(POM).20 They discovered that during times of low POM, the ambient relative humidity and particulate nitrate concentration regulated (N2O5), while during periods of high relative humidity, (N2O5) was controlled by the relative amounts of POM:sulfate, with higher gamma values resulting from lower POM concentrations. In these studies, (N2O5) spanned from <0.01 to

>0.03. Similarly, Brown and coworkers observed a 10× decrease (<0.001- 0.017) in (N2O5) correlated with organic and sulfate loadings during studies in the Northeastern United States.19

6

This summary of recent studies illustrates the significant variability in ambient (N2O5) measurements, and indicates a correlation with a variety of compositional factors. Laboratory studies provide a more constrained measure of the impact of each of these components influencing (N2O5). The laboratory studies detailed below focus on the impact of salts and organics on N2O5 reactive uptake.

1.4.2 Uptake to Inorganic Salts: Laboratory Studies

(N2O5) has been measured on a variety of substrates in the laboratory, with NaCl and seawater providing the most relevant matrices for the marine environment and this dissertation.

Stewart et al. measured uptake to particles generated from both NaCl and artificial seawater at

-3 -2 varying relative humidities (RH), finding that (N2O5) spans a range from 1.5×10 to 1.4×10 for

RH 30-80 %.6 Similar measurements by McNeill and coworkers report a gamma range of

1.5×10-2 to 3.4×10-2 on NaCl and laboratory-generated natural seawater aerosol for RH of

51-71 %.5 These results are in good agreement, however do not account for the range of values measured during ambient studies.

1.4.3 The Impact of Particulate Organics: Laboratory Studies

The presence of organics in aerosol particles can create a physical barrier for N2O5, or hinder its diffusion through the organic phase to the aqueous core. Many groups have performed laboratory study investigations into the effect of a variety of organics on N2O5 reactive uptake

-5 -2 5,8,10,15,23–26 kinetics, finding (N2O5) ranging from <2×10 to >1.5×10 . Folkers et al. found

(N2O5) was reduced by a factor of 35 when sodium sulfate particles were coated with 93% volume fraction a-pinene oxidation products, and a factor of 45 reduction when the particles were comprised only of the oxidation product, indicating the reduction in (N2O5) is due to the presence of an organic film.10 Escorcia and coworkers coated ammonium bisulfate particles with

-pinene oxidation products, increasing the organic mass of the particles up to 79 %.15 Their

7

results show an order of magnitude decrease in (N2O5) when 9 % of the particle mass was organic, and smaller decreases following this, suggesting that the presence of an organic layer on the particle either reduces the mass accommodation coefficient or hinders the dissolution and diffusion of N2O5 into the particle bulk following accommodation. Cosman and coworkers looked into the effect of straight and branched surfactants, finding that upon coating H2SO4 particles with straight chain organics of increasing length, a 17-61× reduction in (N2O5) was observed as

9,24 compared to uptake to uncoated H2SO4 particles. However, when phytanic acid, a branched organic surfactant, was used to coat the particles, no suppression in (N2O5) was observed, implying that straight chain organics pack efficiently on particle surfaces, preventing N2O5 from reaching the sub-phase, whereas branched organics are unable to pack in a tight layer, thus

9,24 allowing for N2O5 diffusion.

1.4.4 Laboratory and Field Measurement Agreement

Despite these ambient and laboratory studies, for complex particles such as sea spray aerosol (SSA), the factors controlling (N2O5) remain largely unclear. Ocean particles contain a complex mix of inorganic and organic species stemming from ocean biology, however the identity and subsequent impact of organic species that transition from the sea surface to SSA, which can account for 17- 83 % of sub-micrometer aerosol mass,27–29 are not well constrained and remain an area of current research.

19,21 It is not understood whether the variability seen in coastal (N2O5) measurements is due to nascent SSA compositional variability or due to processing of particles post-production.

While ambient studies have investigated reactive uptake in coastal marine environments,19,21 it is impossible to isolate ocean-generated particles from other aerosol in the atmosphere such as dust, biomass burning particles etc. Laboratory studies often utilize NaCl or artificial sea salt matrices and coat the particles with organic proxies, such as sodium dodecyl sulfate,5 phytanic acid,9,24 and

8 oxidation products of biogenic gases10,15 however, without a baseline ambient measurement of pure SSA, it is challenging to elucidate the forces driving (N2O5) variability and magnitude.

Additionally, many of the laboratory proxies used are not representative of the natural organics present in ocean water.

1.5 Product Formation Following N2O5 Reaction on Chloride-Containing Surfaces

As mentioned above, when N2O5 dissociates within a particle containing a halide ion, a nitryl halide species is formed. In the marine environment, particles of oceanic origin contain Cl-,

8,16–18 and thus nitryl chloride (ClNO2) is produced following N2O5 uptake. ClNO2 is initially produced in the aqueous phase (R1.4), and with a Henry's law constant of 2.4 - 4.6×10-2 mol L-

1 -1 17,30 atm , efficiently partitions out into the gas phase. In the presence of sunlight, ClNO2 undergoes photolysis, with a lifetime of only a few hours after sunrise, resulting in the formation of chlorine radicals and recycling NOx back into the atmosphere (Reaction 1.6).31,32

ClNO2(g) + h  Cl(g) + NO2(g) (R1.6)

Unlike Reaction 3, by proceeding through this reaction pathway, N2O5 reactive uptake is no longer providing a terminal sink for NOx species, but rather regenerating NO2 in the atmosphere and creating highly oxidizing chlorine radicals.31 In fact, the estimated global contribution of Cl to the atmosphere from this reaction alone is 8-22 Tg yr-1.31

The amount of ClNO2 produced following reactive uptake of N2O5 to a particle surface is determined by calculating the yield of ClNO2, (ClNO2) given by Equation 1.2,

(E1.2)

9

where the yield of ClNO2 is given by the change in the amount of ClNO2 relative to the change in the amount of N2O5 present. Laboratory studies have shown (ClNO2) is directly related to the

12,17,33 amount of chloride present in a particle. Roberts et al. noted (ClNO2) ranges between 0.2-

0.8 for NaCl particles with [Cl-] = 0.02-0.5 M,12 while Bertram and Thornton measured the yield

- 33 from mixed NaNO3/NaCl particles as approaching 1 for [Cl ] >1 M. Figure 1.3 summarizes a number of laboratory measurements of (ClNO2) on chloride-containing particles. It is important to note that these studies generally agree that a 0.5 M Cl- concentration (the concentration of chloride in the ocean) produces a yield of 0.8. Though these studies are in agreement, ambient particles are complex mixtures of chemical components, and thus show a wide variability in

(ClNO2) upon reaction with N2O5.

- Figure 1.3 ClNO2 yield following N2O5 reaction as a function of [Cl ] in solution from which particles were atomized. The results shown are from 3 separate studies.

1.5.1 Ambient Measurements of ClNO2 and (ClNO2)

In 2008, Osthoff and coworkers reported the first ambient observation of ClNO2, finding that in the Gulf of Mexico marine boundary layer, ClNO2 correlated with N2O5 at night, and was

10 produced in abundances up to 1 ppb.34 However, in an urban environment, Mielke et al. found peak ClNO2 levels only reached 0.25 ppb, indicating that reaction with particulate chloride was a

35 minor product of N2O5 uptake compared to heterogeneous hydrolysis. Additionally, ship-based measurements sailing up the coast of California resulted in a maximum [ClNO2] = 2.1 ppb when the air mass was influenced by Los Angeles outflow,36 yet ship-based measurements in the Long

Island Sound resulted in average ClNO2 measurements between 150-250 ppt, while hundreds of miles from the coast in the North Atlantic, levels were between 10-100 ppt.37 A recent study measured vertical profiles of ClNO2 and N2O5 along with particle composition analysis in a continental region. During two pollution plume events, they found ClNO2 levels reach 0.6 and 1.3 ppb and estimated (ClNO2) between 0.3-1, almost the entire allowable range, stating the reason for this large range is uncertain.

1.6 Summary and Remaining Areas to Address

Understanding the driving forces of (N2O5) variability and constraining (N2O5) in the atmosphere is important, as the value used in models to predict NOx removal at night is highly

20 dependent on the (N2O5) value used, as is shown in Figure 1.4 adapted from Bertram et al.

Here it is evident that the choice of (N2O5) not only impacts NOx removal from 0 to 50 % for gamma values ranging from 10-5 to 10-1 , but also has a profound influence on chlorine radical production following ClNO2 generation and photolysis. This has important implications for regional climate models that include heterogeneous chemical reactions in their predications of atmospheric oxidant loadings and air quality.

11

Figure 1.4 Model calculations illustrating the dependence of NOx loss (left y axis) and chlorine radical production following photolysis of ClNO2 (right y axis) on the magnitude of g(N2O5). This plot has been adapted from Bertram et al. 2009.20

As is illustrated above, laboratory and ambient measurements span almost the entire (N2O5) range shown in Figure 1.4, thus constraining our understanding particulate compositional drivers on (N2O5) remains necessary.

Despite the strives made in this area and the detailed investigations into (N2O5) and

(ClNO2), critical areas still remain to be addressed:

1) What is the role of particle mixing state on (N2O5)? Many model parameterizations and heterogeneous reaction kinetics calculations assume particles are homogeneous within a population, meaning they all exhibit some average chemical composition. However, in the ambient atmosphere, particles are not necessarily of an average composition, especially near large aerosol sources, such as the ocean. The impact of mixing state and the effect of average chemical composition assumptions has yet to be addressed.

2) What is (N2O5) on fresh nascent sea spray that is unperturbed by oxidants and other atmospheric influences? Further, how do biological conditions in the ocean impact (N2O5) by altering the chemical composition of SSA, and do the organics present in SSA act as efficient

12

surfactants, pack tightly, and reduce (N2O5)? To date the literature lacks direct studies of (N2O5) on nascent sea spray aerosol that allow for assessment of the role that organics on the surface of

SSA may play in N2O5 uptake kinetics.

3) How do organics present in the top layers of the ocean impact the ability of the ocean to act as an efficient source of ClNO2 following reaction with N2O5? Additionally, are there inorganic mechanisms that might influence (ClNO2) from the ocean surface?

1.7 Synopsis of Chapters

This dissertation addresses the disconnect in ambient and laboratory measurements of

(N2O5) and (ClNO2), utilizing ambient field data, laboratory measurements, real ocean water systems and representative sea spray generation mechanisms. Specifically this dissertation seeks to investigate the three unaddressed areas listed above.

Chapter 2 details investigations into the role of aerosol particle mixing in the marine boundary layer on (N2O5) by correlating (N2O5) measurements made in a polluted coastal environment with chloride-to-nitrate particulate loadings in the first simultaneous (N2O5) ambient observations with co-located single particle chemical composition. This work compares model predictions of (N2O5) in both polluted and clean coastal conditions, suggests controlling factors, and provides a framework for assessment of the role of mixing state in N2O5 reactive uptake to complex ambient aerosol.

In Chapter 3, the role of biological ocean conditions is investigated in a laboratory study, specifically probing the effect of organics during a controlled synthetic mesocosm and the impact they have on SSA and (N2O5). The organic volume fractions of the particles generated from the synthetic mesocosm are measured, and the data is correlated to the observed (N2O5) values.

Additionally, real ocean water is used to generate nascent SSA and the associated (N2O5) values are used for comparison with synthetic mesocosm-derived gammas.

13

Chapters 4 and 5 provide laboratory measurements of laboratory proxy and natural seawater from multiple locations and during a phytoplankton bloom to constrain the role of organics (chapter 4) and inorganics (chapter 5) in controlling the production of ClNO2 from the ocean surface, which was found in previous literature, to be minimal. The organic investigation utilizes the ratio of reactive organic species to chloride at the air-water interface to illustrate how

+ organic enhancement at the ocean surface could be responsible for titrating NO2 , resulting in low to minimal (ClNO2). Chapter 5 similarly investigates the role of inorganic ions in the reduction

2+ of ClNO2 production from the ocean surface, focusing on the Mg effects.

1.8 References

(1) Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd Edition; Wiley-Interscience, 2006.

(2) Finlayson-Pitts, B. J.; Jr, J. N. P. Chemistry of the Upper and Lower Atmosphere: Theory, Experiments, and Applications; Academic Press, 1999.

(3) Mozurkewich, M.; Calvert, J. G. Reaction Probability of N2O5 on Aqueous Aerosols. J. Geophys. Res. Atmospheres 1988, 93 (D12), 15889–15896.

(4) Morris, E. D.; Niki, H. Reaction of Dinitrogen Pentoxide with Water. J. Phys. Chem. 1973, 77 (16), 1929–1932.

(5) McNeill, V. F.; Patterson, J.; Wolfe, G. M.; Thornton, J. A. The Effect of Varying Levels of Surfactant on the Reactive Uptake of N2O5 to Aqueous Aerosol. Atmos Chem Phys 2006, 6 (6), 1635–1644.

(6) Stewart, D. J.; Griffiths, P. T.; Cox, R. A. Reactive Uptake Coefficients for Heterogeneous Reaction of N2O5 with Submicron Aerosols of NaCl and Natural Sea Salt. Atmos Chem Phys 2004, 4 (5), 1381–1388.

(7) Thornton, J. A.; Braban, C. F.; Abbatt, J. P. D. N2O5 Hydrolysis on Sub-Micron Organic Aerosols: The Effect of Relative Humidity, Particle Phase, and Particle Size. Phys. Chem. Chem. Phys. 2003, 5 (20), 4593–4603.

(8) Thornton, J. A.; Abbatt, J. P. D. N2O5 Reaction on Submicron Sea Salt Aerosol: Kinetics, Products, and the Effect of Surface Active Organics. J. Phys. Chem. A 2005, 109 (44), 10004–10012.

14

(9) Cosman, L. M.; Bertram, A. K. Reactive Uptake of N2O5 on Aqueous H2SO4 Solutions Coated with 1-Component and 2-Component Monolayers. J. Phys. Chem. A 2008, 112 (20), 4625–4635.

(10) Folkers, M.; Mentel, T. F.; Wahner, A. Influence of an Organic Coating on the Reactivity of Aqueous Aerosols Probed by the Heterogeneous Hydrolysis of N2O5. Geophys. Res. Lett. 2003, 30 (12), 1644.

(11) Griffiths, P. T.; Badger, C. L.; Cox, R. A.; Folkers, M.; Henk, H. H.; Mentel, T. F. Reactive Uptake of N2O5 by Aerosols Containing Dicarboxylic Acids. Effect of Particle Phase, Composition, and Nitrate Content. J. Phys. Chem. A 2009, 113 (17), 5082–5090.

(12) Roberts, J. M.; Osthoff, H. D.; Brown, S. S.; Ravishankara, A. R.; Coffman, D.; Quinn, P.; Bates, T. Laboratory Studies of Products of N2O5 Uptake on Cl− Containing Substrates. Geophys. Res. Lett. 2009, 36 (20), L20808.

(13) Wahner, A.; Mentel, T. F.; Sohn, M.; Stier, J. Heterogeneous Reaction of N2O5 on Sodium Nitrate Aerosol. J. Geophys. Res. Atmospheres 1998, 103 (D23), 31103–31112.

(14) Mentel, T. F.; Sohn, M.; Wahner, A. Nitrate Effect in the Heterogeneous Hydrolysis of Dinitrogen Pentoxide on Aqueous Aerosols. Phys. Chem. Chem. Phys. 1999, 1 (24), 5451– 5457.

(15) Escorcia, E. N.; Sjostedt, S. J.; Abbatt, J. P. D. Kinetics of N2O5 Hydrolysis on Secondary Organic Aerosol and Mixed Ammonium Bisulfate−Secondary Organic Aerosol Particles. J. Phys. Chem. A 2010, 114 (50), 13113–13121.

(16) Brown, S. S.; Ryerson, T. B.; Wollny, A. G.; Brock, C. A.; Peltier, R.; Sullivan, A. P.; Weber, R. J.; Dubé, W. P.; Trainer, M.; Meagher, J. F.; Fehsenfeld, F. C.; Ravishankara, A. R. Variability in Nocturnal Nitrogen Oxide Processing and Its Role in Regional Air Quality. Science 2006, 311 (5757), 67–70.

(17) Bertram, T. H.; Thornton, J. A.; Riedel, T. P.; Middlebrook, A. M.; Bahreini, R.; Bates, T. S.; Quinn, P. K.; Coffman, D. J. Direct Observations of N2O5 Reactivity on Ambient Aerosol Particles. Geophys. Res. Lett. 2009, 36 (19), L19803.

(18) Finlayson-Pitts, B. J.; Ezell, M. J.; Pitts, J. N. Formation of Chemically Active Chlorine Compounds by Reactions of Atmospheric NaCl Particles with Gaseous N2O5 and ClONO2. Nature 1989, 337 (6204), 241–244.

(19) Behnke, W.; George, C.; Scheer, V.; Zetzsch, C. Production and Decay of ClNO2 from the Reaction of Gaseous N2O5 with NaCl Solution: Bulk and Aerosol Experiments. J. Geophys. Res. Atmospheres 1997, 102 (D3), 3795–3804. (20) Francis Schweitzer, P. M. Multiphase Chemistry of N2O5, ClNO2, and BrNO2. J. Phys. Chem. A 1998, 102 (22).

(21) Riedel, T. P.; Bertram, T. H.; Ryder, O. S.; Liu, S.; Day, D. A.; Russell, L. M.; Gaston, C. J.; Prather, K. A.; Thornton, J. A. Direct N2O5 Reactivity Measurements at a Polluted Coastal Site. Atmos Chem Phys 2012, 12 (6), 2959–2968.

15

(22) Wagner, N. L.; Riedel, T. P.; Young, C. J.; Bahreini, R.; Brock, C. A.; Dubé, W. P.; Kim, S.; Middlebrook, A. M.; Öztürk, F.; Roberts, J. M.; Russo, R.; Sive, B.; Swarthout, R.; Thornton, J. A.; VandenBoer, T. C.; Zhou, Y.; Brown, S. S. N2O5 Uptake Coefficients and Nocturnal NO2 Removal Rates Determined from Ambient Wintertime Measurements. J. Geophys. Res. Atmospheres 2013, 118 (16), 9331–9350.

(23) Anttila, T.; Kiendler-Scharr, A.; Tillmann, R.; Mentel, T. F. On the Reactive Uptake of Gaseous Compounds by Organic-Coated Aqueous Aerosols: Theoretical Analysis and Application to the Heterogeneous Hydrolysis of N2O5. J. Phys. Chem. A 2006, 110 (35), 10435–10443.

(24) Cosman, L. M.; Knopf, D. A.; Bertram, A. K. N2O5 Reactive Uptake on Aqueous Sulfuric Acid Solutions Coated with Branched and Straight-Chain Insoluble Organic Surfactants. J. Phys. Chem. A 2008, 112 (11), 2386–2396.

(25) Knopf, D. A.; Cosman, L. M.; Mousavi, P.; Mokamati, S.; Bertram, A. K. A Novel Flow Reactor for Studying Reactions on Liquid Surfaces Coated by Organic Monolayers: Methods, Validation, and Initial Results. J. Phys. Chem. A 2007, 111 (43), 11021–11032.

(26) Park, S.-C.; Burden, D. K.; Nathanson, G. M. The Inhibition of N2O5 Hydrolysis in Sulfuric Acid by 1-Butanol and 1-Hexanol Surfactant Coatings. J. Phys. Chem. A 2007, 111 (15), 2921–2929.

(27) Facchini, M. C.; Rinaldi, M.; Decesari, S.; Carbone, C.; Finessi, E.; Mircea, M.; Fuzzi, S.; Ceburnis, D.; Flanagan, R.; Nilsson, E. D.; de Leeuw, G.; Martino, M.; Woeltjen, J.; O’Dowd, C. D. Primary Submicron Marine Aerosol Dominated by Insoluble Organic Colloids and Aggregates. Geophys. Res. Lett. 2008, 35 (17), L17814.

(28) O’Dowd, C. D.; Facchini, M. C.; Cavalli, F.; Ceburnis, D.; Mircea, M.; Decesari, S.; Fuzzi, S.; Yoon, Y. J.; Putaud, J.-P. Biogenically Driven Organic Contribution to Marine Aerosol. Nature 2004, 431 (7009), 676–680.

(29) Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; DeMott, P. J.; Aluwihare, L. I.; Palenik, B. P.; Azam, F.; Seinfeld, J. H.; Moffet, R. C.; Molina, M. J.; Cappa, C. D.; Geiger, F. M.; Roberts, G. C.; Russell, L. M.; Ault, A. P.; Baltrusaitis, J.; Collins, D. B.; Corrigan, C. E.; Cuadra-Rodriguez, L. A.; Ebben, C. J.; Forestieri, S. D.; Guasco, T. L.; Hersey, S. P.; Kim, M. J.; Lambert, W. F.; Modini, R. L.; Mui, W.; Pedler, B. E.; Ruppel, M. J.; Ryder, O. S.; Schoepp, N. G.; Sullivan, R. C.; Zhao, D. Bringing the Ocean into the Laboratory to Probe the Chemical Complexity of Sea Spray Aerosol. Proc. Natl. Acad. Sci. 2013, 110 (19), 7550–7555.

(30) Frenzel, A.; Scheer, V.; Sikorski, R.; George, C.; Behnke, W.; Zetzsch, C. Heterogeneous Interconversion Reactions of BrNO2, ClNO2, Br2, and Cl2. J. Phys. Chem. A 1998, 102 (8), 1329–1337.

(31) Thornton, J. A.; Kercher, J. P.; Riedel, T. P.; Wagner, N. L.; Cozic, J.; Holloway, J. S.; Dubé, W. P.; Wolfe, G. M.; Quinn, P. K.; Middlebrook, A. M.; Alexander, B.; Brown, S.

16

S. A Large Atomic Chlorine Source Inferred from Mid-Continental Reactive Nitrogen Chemistry. Nature 2010, 464 (7286), 271–274.

(32) Nelson, H. H.; Johnston, H. S. Kinetics of the Reaction of Chlorine Atoms with Nitrosyl Chloride and Nitryl Chloride and the Photochemistry of Nitryl Chloride. J. Phys. Chem. 1981, 85 (25), 3891–3896.

(33) Bertram, T. H.; Thornton, J. A. Toward a General Parameterization of N2O5 Reactivity on Aqueous Particles: The Competing Effects of Particle Liquid Water, Nitrate and Chloride. Atmos Chem Phys 2009, 9 (21), 8351–8363.

(34) Osthoff, H. D.; Roberts, J. M.; Ravishankara, A. R.; Williams, E. J.; Lerner, B. M.; Sommariva, R.; Bates, T. S.; Coffman, D.; Quinn, P. K.; Dibb, J. E.; Stark, H.; Burkholder, J. B.; Talukdar, R. K.; Meagher, J.; Fehsenfeld, F. C.; Brown, S. S. High Levels of Nitryl Chloride in the Polluted Subtropical Marine Boundary Layer. Nat. Geosci. 2008, 1 (5), 324–328.

(35) Mielke, L. H.; Furgeson, A.; Osthoff, H. D. Observation of ClNO2 in a Mid-Continental Urban Environment. Environ. Sci. Technol. 2011, 45 (20), 8889–8896.

(36) Riedel, T. P.; Bertram, T. H.; Crisp, T. A.; Williams, E. J.; Lerner, B. M.; Vlasenko, A.; Li, S.-M.; Gilman, J.; de Gouw, J.; Bon, D. M.; Wagner, N. L.; Brown, S. S.; Thornton, J. A. Nitryl Chloride and Molecular Chlorine in the Coastal Marine Boundary Layer. Environ. Sci. Technol. 2012, 46 (19), 10463–10470.

(37) Kercher, J. P.; Riedel, T. P.; Thornton, J. A. Chlorine Activation by N2O5: Simultaneous, in Situ Detection of ClNO2 and N2O5 by Chemical Ionization Mass Spectrometry. Atmos Meas Tech 2009, 2 (1), 193–204.

Chapter 2 On The Role of Particle Inorganic Mixing State in The Reactive Uptake of N2O5 to Ambient Aerosol Particles

2.1 Abstract

The rates of heterogeneous reactions of trace gases with aerosol particles are complex functions of particle chemical composition, morphology, and phase state. Currently, the majority of model parameterizations of heterogeneous reaction kinetics focus on the population average of aerosol particle mass, assuming individual particles have the same chemical composition as the average state. Here, we assess the impact of particle mixing state on heterogeneous reaction kinetics using the N2O5 reactive uptake coefficient, (N2O5), dependence on the particulate

- - chloride-to-nitrate ratio (nCl /nNO3 ). We describe the first simultaneous ambient observations of single particle chemical composition and in situ determinations of (N2O5). When accounting for

- - particulate nCl /nNO3 mixing state, model parameterizations of (N2O5) continue to over predict

(N2O5) by more than a factor of two in polluted coastal regions, suggesting that chemical composition and physical phase state of particulate organics likely control (N2O5) in these air masses. In contrast, direct measurement of (N2O5) in air masses of marine origin are well captured by model parameterizations and reveal limited suppression of (N2O5), indicating the organic mass fraction of fresh sea spray aerosol at this location does not suppress (N2O5). We provide an observation-based framework for assessing the impact of particle mixing state on gas- particle interactions.

2.2 Introduction

The production rates of tropospheric ozone (O3) and secondary organic aerosol (SOA) are dependent on the abundance of nitrogen oxides (NOx), the concentration and speciation of volatile organic compounds (VOC), and oxidant loadings (e.g., OH and Cl). Accurate assessment

17

18

of the dependence of O3 and SOA production on NOx emission rates requires a detailed description of NOx removal mechanisms and the coupling of these processes with other oxidants.

Laboratory measurements, confirmed by field observations, have shown that the heterogeneous reaction of dinitrogen pentoxide (N2O5) with aerosol particles serves as an efficient loss mechanism for NOx, where the fraction of emitted NOx that is removed from the atmosphere via this pathway is estimated to be as large as 50% by current chemical transport models.1 Further, the reaction of N2O5 on chloride containing particles results in the production of nitryl chloride

(ClNO2), that has been estimated to account for as much as 60% of the total chlorine radical

2 production rate in polluted coastal regions. Despite both the importance of N2O5 chemistry to oxidant loadings and an abundance of detailed mechanistic studies of the reaction mechanism on single-component particle surrogates in the laboratory, accurate parameterization of N2O5 reactive uptake in the atmosphere remains challenging.3

2.2.1 Factors Controlling the Reactive Uptake of N2O5 to Aerosol Particles

1,4 Direct measurement of the reactive uptake coefficient, (N2O5), to ambient aerosol, and determinations of (N2O5) through steady-state analysis indicate that parameterizations do not

3,5 capture the observed magnitude and variability in (N2O5). Model-measurement disagreement is most pronounced in regions characterized by large chemical heterogeneity within the particle

3 population, where parameterizations routinely overestimate (N2O5). This disagreement arises from the dependence of (N2O5) on single particle chemical composition, physical phase state, morphology, and liquid water content.

Several key components of the N2O5 reaction mechanism have been defined: 1) the presence of organic films/coatings6,7 and monolayer surfactants8–11 has been shown to suppress

12 + the rate of N2O5 uptake and 2) the competition for the N2O5 dissociation product NO2 , between

- - strong nucleophiles (e.g., H2O and Cl ) and the nitrate ion (NO3 ), can act to accelerate or suppress

19

13–15 (N2O5), respectively. Both competition processes highlight how the distribution of chemicals within a single particle or particle population can impact reaction mechanisms, as the reaction rate depends on the single particle composition, not the ensemble average concentrations.

2.2.2 Single Particle Chloride-to-Nitrate Mixing State in The Polluted Marine Boundary

Layer

The chemical compositions of individual particles within a population of particles have varying degrees of chemical similarity to the population average. In one extreme, all particles are chemically identical to the population average (internally mixed), and in the other extreme, individual particles, or size regimes of particles, have unique chemical signatures, which differ from the mean state (externally mixed).16 The extent to which particles are internally or externally mixed within any given air mass is a function of particle source and chemical and physical processing (i.e. heterogeneous reactions, condensation, etc.) post-emission or formation.

Typically, particle populations are externally mixed near source regions, where primary aerosol particles are emitted to the atmosphere with characteristic size and chemical composition unique to the source mechanism.17,18 Within a Lagrangian framework, as the air mass evolves in time away from the source region, it is thought that the particle population will tend toward an internal mixture as: 1) the semi-volatile components of the primary particles redistribute themselves among the entire population,19 2) low-volatility products of gas-phase reactions condense onto existing surfaces20 and 3) coagulation and deposition act to consolidate particles into the

18,21 accumulation mode (0.08 µm < particle diameter (dp) < 2.5 µm). While a trajectory toward an internally mixed state is consistent with a particle population dominated by semi-volatile organic compounds, particle morphology and physical phase state could act to slow the approach toward an internally mixed state, and the presence of organic and inorganic acids and bases can result in pH differences between particles of different sizes and limit the formation of an internally mixed

20 population.22,23 Further, a particle population within a stagnant fixed coordinate system, where particles of specific composition are continuously added, such as in the marine boundary layer

(MBL), can remain as an external mixture when there is a kinetic or diffusion limitation to gas- particle exchange, as particles within the population are of different ages.

Heterogeneous and multiphase reactions occurring on or within sea spray aerosol are of particular interest in atmospheric chemistry as they can serve as halogen activation mechanisms.

Here, we discuss the activation of particulate chloride to ClNO2 which rapidly photolyzes to produce Cl radicals. Nascent sea spray aerosol (SSA) is heavily enriched in Cl- with respect to

– - - 4 16 NO3 due to the high molar chloride-to-nitrate ratio (nCl /nNO3 ) of seawater ( > 3×10 ). The chemical processing of SSA by nitric acid (HNO3) and N2O5 in the polluted MBL results in rapid

- - - - depletion of particulate Cl and accumulation of NO3 until nCl /nNO3 achieves equilibrium with gas-phase HNO3 and HCl (R1-R2).

  Cl aq HNO3 gHCl g  NO3 aq (R1)

  Cl aq  N2O5 g ClNO2 g  NO3 aq (R2)

Given the volatility of both HCl and HNO3, and that each particle is exposed to the same

- - HCl/HNO3 gas-phase ratio, it is expected that nCl /nNO3 would be internally mixed within the particle population and in equilibrium with the gas-phase. However, concurrent field observations of particle and gas-phase nitrate and chloride in the polluted boundary layer have shown nCl-

- /nNO3 is often larger on super-micrometer particles as compared with sub-micrometer particles and the super-micrometer component is often not in equilibrium with the gas-phase. 22,24–26 This result suggests that either: 1) inter-particle variability in pH or liquid water content acts to sustain

- - particle-particle variability in nCl /nNO3 , 2) the approach to equilibrium is slowed due to the

21 presence of organic films, 3) fast production and size-dependent deposition compete on the time- scale required to establish equilibrium, or 4) HNO3 reactions on or within super-micrometer particles are diffusion-limited.27

- - In what follows, we use direct measurements of single particle nCl /nNO3 and ambient

- - determinations of (N2O5) to assess the impact of nCl /nNO3 mixing state on (N2O5) and the potential use for single particle measurements to better constrain parameterizations of (N2O5).

The results obtained here provide an upper limit on the role of inorganic mixing state in determining (N2O5), where any remaining differences between observation and parameterization likely result from the direct impact of organic coatings and/or their control over particle liquid

H2O.

2.3 Materials and Methods

Ambient measurements of single particle chemical composition and in situ

1 determinations of N2O5 reactivity were conducted simultaneously between October 4th-6th 2009 from the Scripps Institution of Oceanography (SIO) Pier. The flow reactor apparatus and

28 modulation technique implemented for determining (N2O5) have been described previously.

Briefly, ambient air was sampled through a 0.25" ID stainless steel inlet either directly into the flow reactor, or via a Teflon filter, providing 99% particle removal from the air stream, into a 5.9"

ID stainless steel tube, 35" in length. N2O5 was produced in situ via the reaction of excess NO2 with O3, created via photolysis of O2, in a secondary mixing cell. N2O5 was then injected into the reactor, allowing for reaction of the gas with ambient aerosol. The mixing ratio of N2O5 at the top of the flow reactor was ca. 1 ppbv, and was monitored using a chemical ionization mass spectrometer. In this application, an Aerosol Time of Flight Mass Spectrometer (ATOFMS)29 sampled both ambient aerosol as well as ambient aerosol that had been reacted with N2O5 permitting analysis of the extent of chloride depletion on a single particle level.

22

2.4 Results and Discussion

2.4.1 Impact of Particulate Chloride-to-Nitrate Mixing State on (N2O5)

The impact of particle mixing state on heterogeneous and multi-phase reactions is most pronounced when the property of interest is a nonlinear function of the ratio of two chemical components, resulting in different ensemble average kinetics for the same average chemical

30 composition. Recently, Bertram and Thornton suggested that (N2O5), in the absence of organic coatings, could be parameterized as:

   

'  1   N O  Ak2 f 1  (E1) 2 5  k H O(l)  k Cl      3 2   1  4             k2b NO3    k2b NO3  

+ - where the aqueous phase reactions are described by k2f for N2O5 + H2O, k2b for NO2 + NO3 , k3

+ + for NO2 + H2O, and k4 for a halide with NO2 . Additionally, A is an empirical pre-factor that includes V (the total particle volume concentration), Sa (the total particle surface area concentration), the mean molecular speed of  and KH (the dimensionless Henry’s law coefficient (KH ≡[N2O5]aq /[N2O5]g )). The parameterization highlights the dependence of (N2O5)

- - - on both particle nH2O/nNO3 and nCl /nNO3 , and suggests that (N2O5) will respond non-linearly

- - - - to changes in nCl /nNO3 and that individual particles with high nCl /nNO3 will be more reactive.

A hypothetical graphical representation of the (N2O5) parameterization is shown in Fig.

- - - 2.1A, where nascent SSA are emitted at high nH2O/nNO3 and nCl /nNO3 (red area/upper right), and exhibit large (N2O5) in the absence of organic films. As nascent SSA is chemically

- - processed in the atmosphere, nCl /nNO3 decreases and the rate of N2O5 uptake slows (lower left).

23

The impact of the particle mixing state assumption on the parameterized (N2O5) is shown in Fig.

2.1B, using the probability density function (PDF) for two hypothetical distributions. The PDF is employed as a means to examine both the extent of particle aging between populations and the degree of external mixing within a specific particle population. Calculation of (N2O5) for the particle population shown in Fig. 2.1B, can be conducted either: 1) on a single particle basis where E1 is applied to each individual particle in the population and the ensemble average

(N2O5) then calculated as the average of the single particle determinations (referred to here as

- - the externally mixed model) or 2) using the population ensemble average nCl /nNO3 (referred to here as the internally mixed model) to calculate the (N2O5). Comparison of these two model parameterizations to measured ambient values exposes the possible error that can result from assuming a particle population is internally mixed.

Calculation of (N2O5) for the externally mixed population shown in Fig. 2.1B, yields a

- population averaged reactive uptake coefficient of 0.025 (for nH2O/nNO3 = 3). In contrast,

- - - - assuming complete internal mixing of nCl /nNO3 (with same ensemble average nCl /nNO3 = 0.1) results in a 32% overestimation of (N2O5). The schematic representation highlights the potential importance for resolving particle mixing state, as gas-particle interactions depend on the surface and bulk chemical composition of single particles, which can be different from the ensemble average.

24

- Figure 2.1 A) Model parameterization of (N2O5), where nascent SSA are emitted at high nH2O/nNO3 (aq) - - and nCl /nNO3 , and exhibit large (N2O5) in the absence of organic films. As nascent SSA is chemically - - processed in the atmosphere, nCl /nNO3 decreases and the rate of N2O5 uptake slows. B) The impact of the particle mixing state assumption on the parameterized (N2O5) for two hypothetical distributions with the - - same ensemble average nCl /nNO3 . In this example, the internal mixing assumption results in a 32% overestimation of (N2O5)

In what follows, we use ambient measurements of single particle chemical composition retrieved from the ATOFMS as an input for the aforementioned (N2O5) parameterization. This permits calculation of (N2O5) under both internal and external mixing assumptions for

- comparison to the measured ambient value. The value of nH2O/nNO3 used in this analysis was determined for each case via the Aerosol Inorganics Model31, using measurements of ambient RH

- and ATOFMS derived chemical composition as input parameters. Values for nH2O/nNO3 ranged from 3.5 to greater than 6000. The parameterization of N2O5) under these conditions is

- relatively insensitive to the absolute value of nH2O/nNO3 . Error analysis is discussed in detail in

25 the supporting online information, however two important details regarding the application of the

ATOFMS data and the (N2O5) parameterization are as follows: 1) the parameterization limits (N2O5) to ca. 0.04, thus bounding calculation of the upper limit uncertainty, and 2) the

- - - - ATOFMS-measured single particle nCl /nNO3 using standard solutions of known Cl /NO3 is biased low, resulting in a reduced gamma value by at most 7% for the reported ambient pier measurements.

Figure 2.2 A) Backward air trajectories calculated for 24 hours prior to sampling for Period 1 and Period 2. B) Particle number concentration (N, left axis; see supporting information for sizing measurement details) - - and NOx (right axis) and C) Median nCl /nNO3 as determined by ATOFMS over the course of the two sampling periods (grey) with interquartile range (grey shaded region).

26

2.4.2 Observation-Based Constraints on the Impact of Chloride-to-Nitrate Particle

Mixing State on (N2O5)

We focus our ambient analysis on two specific air masses sampled from the Scripps Pier site during Fall 2009. To better understand the history of the air masses, HYSPLIT backwards trajectories were calculated for 24 hours prior to the sampling period at 500m above sea level using GDAS archived meteorological data.32,33 As shown in Fig. 2.2A, back trajectories calculated for Period 1 (October 3rd) were characterized by stagnant winds, where particles sampled from the SIO pier had significant interaction with coastal pollution. During this period, average relative humidity, measured at an adjacent coastal site, was 71%, NOx mixing ratios at the site averaged 10 ± 8 ppb, while sub-micrometer particle number concentration averaged 2477

± 23 particles cm-3 (Fig. 2.2B), consistent with previous literature values for a polluted marine boundary layer at this location.34 As expected, the median single particle measurement of nCl-

- /nNO3 (Fig. 2.2C) was 0.15 (interquartile range: 0.046 - 0.47), confirming that particles sampled during this period had undergone significant chemical processing. In contrast, 24 hour back trajectory analysis indicates strong westerly winds for the second sampling period (Period 2;

October 4th). As such, Period 2 was characterized by an average relative humidity of 64%, significantly lower NOx mixing ratios, 0.92 ± 0.5 ppb, particle concentrations, 864 ± 4 particles

-3 - - cm , and higher particulate nCl /nNO3 (0.63, interquartile range: 0.21 - 3.0) indicative of a less- polluted marine source for the air sampled. These values are summarized in Table 2.1. As

- expected, coincident in situ measurements of N2O5) reflect the average differences in nCl

- /nNO3 , where N2O5) was measured to be 0.012 ± 0.007 (1) for Period 1 (polluted air) and

0.037 ± 0.009 (1) for Period 2 (less-polluted marine air).

For the less-polluted marine conditions (Period 2), calculation of N2O5) using the

Bertram and Thornton30 parameterization (E1), assuming internal mixing, agrees well with the

‡ observed N2O5) (N2O5)internal mixing parameterization = 0.036 , as compared with

27

Table 2.1N2O5 values calculated using the assumption of an externally mixed particle population, an internally mixed population, and measured from SIO pier, along average RH and particle properties for sampling Periods 1 and 2.

N2O5: N2O5: N2O5: Sub- Calculated Calculated Measurement Average micrometer Average Median Average m/Q Location Model; Model; NO Particle f RH x Cl/NO ROA 37/(37+43) Externally Internally (ppb) Conc. 3 3 Mixed Mixed (#/cm )

SIO Pier; 0.15, (iqr= 0.031 + 0.012 0.012 ± 2.69×10-3 more- 0.040‡ 71% 10 ± 8 2477 ± 23 0.046 - 0.448 ± 0.42 / - 0.008* 0.007 ±0.01 polluted 0.47)

SIO Pier; 0.63, (iqr 0.033 + 0.005 0.037 ± 0.92 ± 1.64×10-3 less- 0.036‡ 64% 864 ± 4 = 0.21 - 0.485 ± 0.42 / -0.006* 0.009 0.5 ±0.004 polluted 3.0)

- - *These error values account for both the accuracy and narrow distribution of nCl /nNO3 ‡Calculation of error values for these values results in error bounds that are not statistically different from the average value reported. This is due to the nature of the internally mixed calculation, and narrow - - distribution of nCl /nNO3 within this calculation. The absence of error bounds here however, does not represent the absolute accuracy of the value.

N2O5)meas = 0.037). Model-measurement agreement and N2O5) values consistent with

35 deliquesced NaCl (N2O5) = 0.03 ) suggest a more internally mixed aerosol population and, indirectly indicate, an insignificant role for organic films in suppressing heterogeneous reactions on nascent SSA, under the conditions sampled here. This result suggests that nascent SSA (at the surface area maximum) sampled at the SIO pier do not contain a significant organic film or surfactant coating that suppresses N2O5). While this is counter to previous laboratory suggestions conducted using straight chain organic model systems, 8,35 it is consistent with results obtained on branched organic chain systems. (Note: ‡ These error values indicate the narrow

- - distribution of nCl /nNO3 within the population, however the error does not represent the absolute accuracy of the value).

However, for polluted conditions, the internal mixing parameterization overestimates the

‡ observed N2O5) by a factor of 3 (N2O5); internal mixing parameterization = 0.040 , as

28

compared with N2O5)meas = 0.012). Overestimation of N2O5) for polluted conditions (Period 1)

- - can be attributed to either: 1) external mixing of Cl and NO3 or 2) a significant role for particulate organics in regulating particle liquid water concentration or access to it through the

1 - - formation of an organic film. Here, we use single particle measurements of Cl and NO3 to constrain the role of inorganic mixing state.

As discussed above, N2O5) can be calculated on a single particle basis, and then averaged for comparison with the measured N2O5). Using this approach, we calculate an average N2O5) of 0.031 (+0.012, -0.008), a 17% reduction in N2O5) as compared to the analogous calculation assuming internal mixing. This result suggests that inorganic mixing state is not the major controller in the observed suppression of N2O5), but rather that the organic component is acting to suppress N2O5) due to the amount, type, or phase state of the organic present.

Figure 2.3A illustrates the fraction of particles sampled during Periods 1 and 2 which,

- - based on their nCl /nNO3 values, exhibit a specific range of N2O5) when calculated using the

Bertram and Thornton parameterization under the external mixing assumption. It is evident here, that the fraction of particles producing the given calculated N2O5) ranges for both Period 1

(more-polluted) and Period 2 (less-polluted) are similar, with the least variation between the two for N2O5) > 0.03. This highlights the small differences in reactivity of fractions of the populations during these two sampling periods, as calculated using the parameterization assuming externally mixed particles.

29

Figure 2.3 A) Histogram of calculated single particle (N2O5), based on ATOFMS single particle - - measurements of nCl /nNO3 at the SIO pier, for sampling period 1 (more-polluted; pink) and period 2 - - (less-polluted; blue). B,C) PDF of ATOFMS nCl /nNO3 as measured for sub-micrometer (light line) and super-micrometer (dark line) particles for sampling period 1 and 2, respectively.

The above N2O5) calculation does not account for the correlation between particle

- - 22,24–26 surface area and nCl /nNO3 as has been seen in previous studies. Fig. 2.3B and C depict

- - the PDF of nCl /nNO3 for sub- and super-micrometer aerosol separately, indicating slightly

- - higher nCl /nNO3 for super-micrometer aerosol in the case of Period 2. We calculate N2O5), under the externally mixed population assumption, during Period 1 to be 0.030 (+0.013, -0.008)

30 and 0.031 (+0.012, -0.008), while for Period 2, the values are 0.029 (+0.010, -0.008) and 0.033

(+0.004, -0.005) for sub- and super-micrometer particles, respectively. Given that the super- micrometer surface area accounts for less than 20% of the total surface area under all conditions sampled at the SIO pier, inclusion of surface area weighting does not explain the differences between field observations and model parameterizations at this location.

Figure 2.4 Probability distribution functions (PDF) of (A) fROA (calculated using the representative organic ions, m/Q (37 + 43) / total ion peak area) and (B) the contribution of m/Q 43 to the organic signal (m/Q 43 / m/Q (37 + 43) ) for the more-polluted (red) and less-polluted (blue) pier data sets.

To assess whether notable changes in the organic mass fraction could be deduced from the ATOFMS, we look at the probability distribution functions for the ratio of two organic peak

+ + areas (sum of the 37 and 43 m/Q peaks that represent the C3H and CH3CO fragments) to the total positive ion peak area. Here forward, we refer to this ratio as fROA, or the representative organic mass fraction of single aerosol particles.36 As shown in Fig. 2.4A and summarized in

Table 2.1, there is no discernible difference in fROA for the two sampling periods. We further look

+ at the ratio of oxygenated fragment (43 m/Q, CH3CO ) to the sum of the two organic markers (37

+ 43 m/Q). As shown in Fig. 2.4B, there is again no discernible difference between the two sampling periods. This result suggests that either the molecular composition of the particulate

31 organic species changed between the two sampling periods, or a subtle change in the RH at the sampling site altered the physical phase state of the organics.37,38

2.5 Insight from Single Particle Measurements of Reacted Sea Spray Aerosol

- - In the preceding section, we used ambient measurements of single particle nCl /nNO3 to calculate single particle (N2O5) for comparison with the population average determination of

(N2O5). During the fall 2009 field deployment, we also made observations of single particle

- - chemical composition post-N2O5 reaction using ATOFMS. The PDF of nCl /nNO3 for ambient

- - particles is shown in Fig. 2.5A and 2.5B along with the PDF of nCl /nNO3 for particles having been exposed to N2O5 in the flow reactor for sampling periods 1 and 2, respectively. The PDF of

- - nCl /nNO3 for particles having been exposed to N2O5 during Period 1 (Fig. 2.5A; more-polluted conditions) is statistically indistinguishable from that measured on ambient particles. For

- - comparison, the PDF of nCl /nNO3 for particles having been exposed to N2O5 during Period 2

(Fig. 2.5B; less-polluted conditions) is also statistically indistinguishable from that measured on ambient particles. This indicates that under both conditions, ambient particles underwent very

- - little chemical change with respect to their nCl /nNO3 ratio in the flow reactor. This is due to the experimental conditions of the ambient N2O5 reactivity experiment, where the N2O5 mixing ratio is held to ca. 1 ppbv to limit nitrate accumulation, which would bias the measured reactive uptake coefficient. Under that experimental conditions used here (e.g., [N2O5] = 1 ppbv and average

- - residence time = 258 sec), and a (N2O5) = 0.03, we would expect the nCl /nNO3 ratio at 0.01

(peak of ambient PDF during Period 2) to decrease by <1%. This unfortunately is too small a perturbation to detect via ATOFMS.

32

- - Figure 2.5 Probability distribution function (PDF) as a function of nCl /nNO3 for: A) more-polluted pier - - data, B) less-polluted pier data, and C) nascent SSA data sets. D) Single particle nCl /nNO3 as a function of particulate organic fraction (fROA) for nascent SSA. fROA is calculated using the representative organic ions, m/Q (37 + 43) / total ion peak area. Small blue circles indicate unreacted particles, while small green squares indicate flow tube-reacted particles. Median values are shown in large symbols.

However, direct comparison of the single particle composition of reacted and unreacted particles conducted at higher gas-phase concentrations may provide insight into the properties of aerosol that act to limit reactive uptake (e.g., organic fraction, liquid water content) on a single particle level. To assess this, we explore the use of single particle aerosol measurements of

33

nascent SSA, and nascent SSA reacted with HNO3 at significantly high mixing ratios ([HNO3] =

155 ppbv) to assess the potential for using real-time single particle approaches to probe for the selective reactivity of specific classes of aerosol particles within a population. Details concerning the experimental set up can be found in the supporting online material. Briefly, nascent SSA was generated in a state of the art, 30 m long wave channel located at Scripps Institution of

Oceanography, which was filled with real ocean water pumped through coarse grain filters and subsequently doped with bacteria and phytoplankton cultures to regenerate surface active organic

39 material potentially lost during filtration. Here, nascent SSA was aged via reaction with HNO3 in an entrained aerosol flow reactor at an average exposure time of 114 sec at 155 ppb HNO3. The chemical composition of the particles post-reaction were measured at the base of the flow reactor via ATOFMS. HNO3 was used in place of N2O5 in these experiments due to the larger effective

27 reactive uptake coefficient on nascent SSA ((HNO3) > 0.1) thus enabling significant particle

- - reactions on the short time scale of a flow reactor. The PDF of nCl /nNO3 for nascent SSA

- - particles (blue) is shown in Fig. 2.5C alongside the PDF of nCl /nNO3 for nascent SSA particles having been exposed to HNO3 in the flow reactor (green). The 4-5 order of magnitude shift in the

- - peak nCl /nNO3 indicates that the nascent SSA react significantly in the flow reactor, providing the opportunity to assess the variability in reaction kinetics within the aerosol population.

- - Single particle measurements of the nascent SSA nCl /nNO3 ratio (y-axis) as a function

- - of fROA are shown as blue circles in Fig. 2.5D. As shown, single particle nCl /nNO3 decreases

-6 with increasing fROA, from 686.4 to 4.90 for fROA of 7.81×10 and 0.25, respectively. While the source of this dependence is unknown, it suggests that either: 1) nascent SSA generated in this study is externally mixed displaying more than one type of aerosol particle40 or that 2) ATOFMS is not ionizing the entire particle volume and radial gradients in chloride, nitrate, and organics

- - impact the measurement of nCl /nNO3 at high fROA. For comparison, single particle measurements of nascent SSA having been exposed to HNO3 are also shown, in green squares, in

34

Fig. 2.5D. If all SSA were to react at the same rate, independent of the organic content of the

- - particle, it is expected that nCl /nNO3 would decrease uniformly, independent of fROA. However,

- - in contrast to the nascent SSA, single particle nCl /nNO3 decreases with decreasing fROA, from

-2 -6 76.4 to 2.24×10 for fROA of 0.25 and 7.81×10 , respectively. At high organic content (fROA

- - >0.1), nCl /nNO3 is nearly equal for both the reacted and unreacted particles, indicating that

-6 HNO3 reactive uptake is suppressed. In contrast, at low organic content (fROA = 7.81×10 ),

- - median nCl /nNO3 is over 4 orders of magnitude smaller for the reacted particles, indicating that the particles have undergone significant reaction. This result demonstrates that the presence of organics in nascent SSA, while rare in this study, can act to limit the reactive uptake of HNO3.

Ongoing work is focused on relating fROA quantitatively to the single particle organic mass fraction and potentially an organic film thickness. These results suggest that the reactive uptake of trace gases to nascent SSA may be significantly slower under conditions where the aerosol organic mass fraction has been shown to be large.8,35

2.6 Atmospheric Implications

We present an analysis of the first direct measurements of the dependence of (N2O5) on single particle inorganic mixing state using simultaneous atmospheric observations of (N2O5) and single particle chemical composition. These observations indicate that: 1) when accounting

- - for particulate nCl /nNO3 mixing state, model parameterizations of (N2O5) continue to over predict (N2O5) by more than a factor of two in polluted coastal regions, indicating that the chemical composition and physical phase state of particulate organics likely controls (N2O5) in these air masses, and 2) direct determination of (N2O5) in air masses of marine origin are well captured by model parameterizations and reveal limited suppression of (N2O5) relative to its known value on NaCl particles, indicating that the organic mass fraction of fresh sea spray aerosol at this location does not suppress (N2O5). Surprisingly, simultaneous measurements of

35 single particle organic composition revealed little difference in the mass fraction of organics between the two sampling periods, suggesting that the molecular composition and resulting physical properties of the organic phase of the aerosol control reaction kinetics at this site.

2.7 Acknowledgements

The authors would like to thank Joel Thornton for the use of the CIMS and flow tube apparatus used in the ambient portion of this study. This research was supported by the Office of

Science (Office of Biological and Environmental Research), U.S. Department of Energy (Grant

No. DE-SC0006431) and the National Science Foundation (Grant No. CHE1038028). O.S.R is grateful for a Graduate Research Fellowship from the National Science Foundation (2011- present).

2.8 Supporting Information

2.8.1 Sampling Locations and Details

2.8.1.1 Ambient Study: Scripps Institution of Oceanography Pier

Sampling occurred at the end of Scripps Pier, 300m offshore, located at Scripps

Institution of Oceanography, La Jolla, CA. The simultaneous measurement of N2O5 and single particle data took place between October 4th-6th 2009. Ambient NOx and O3 were measured at the site continuously along with particle size distribution (see below). Relative humidity data was obtained from a meteorological station at a nearby coastal site (<11 miles away) due to absent RH data from the Scripps Pier meteorological station during this time period.

2.8.1.2 Laboratory Study: Nascent Sea Spray Wave Channel

Nascent, fresh sea spray particles were generated in a laboratory setting using a 30 m- long ocean-atmosphere facility, filled with ocean water pumped directly from Scripps Pier after coarse filtration. A hydraulic paddle was utilized to create breaking waves in a fixed location

36 once every second, which accurately reproduced the bubble size distribution created when a wave breaks in the real open ocean.39,41 Nascent SSA discussed here was generated in this fashion from ocean water doped with bacteria (Alteromonas and Pseudoalteromonas atlantica) and phytoplankton (Dunaliella tertiolecta) cultures to reproduce organics possibly lost during coarse filtration. This experimental setup was a large study, employing a suite of instruments, to better understand the factors causing variability in measured SSA properties from ambient field studies.39 Pertinent to this paper, an ATOFMS sampled from the headspace above the breaking waves to allow for measurement of the chemical composition of nascent sea spray. A flow reactor system was used to age the particles with HNO3 prior to measurement, allowing for comparison of pre-and post-reacted particle chemical composition. This flow reactor was a 3" ID Pyrex tube, coated with halocarbon wax to minimize gas and particle interactions with the walls. Particles were injected through a side port at the top of the reactor, and allowed to mix with HNO3 throughout the residence time in the tube. The chemical composition was measured from the base of the reactor. HNO3 was generated using an NO2 permeation tube (Kin-Tek Laboratories, Inc.) coupled to an O3 production cell.

2.8.1.3 Ambient Study: Jeju Island, South Korea

A second ambient data set used as a comparison of single particle chemical composition in this supporting document involves particles sampled from Jeju Island, South Korea.42 Single particle data was collected with an ATOFMS between April 12th-May 15th 2007, with a one week period of data presented in this discussion. The island is located in the East China Sea, where air masses can typically be influenced by outflow from mainland China, South Korea, or fresh marine sources. The sampling site was in a rural location, situated 20 m from the ocean on the west coast of the island to avoid local contamination.

37

2.8.1.4 Ambient Study: Long Beach, California

A third ambient data set used as an example of chemical composition in a highly polluted particle population was sampled in Long Beach, California between November 16-26, 2007.43

ATOFMS measurements were taken from a stationary platform located on Terminal Island in the

Port of Los Angeles, heavily used by ships, allowing for the sampling of both daytime onshore flows, and nighttime offshore flows. For further details on the logistics of this study, please see

Ault et al. 2010.

2.8.2 Measurements

2.8.2.1 N2O5 Reactivity Determinations:

Ambient air from Scripps Pier was sampled through a 0.25" ID stainless steel tube either directly into the flow reactor, or via a filter, providing 99% particle removal from the air stream.

The flow reactor has been described in detail previously.28 Briefly, it is a 5.9" ID stainless steel tube mounted vertically, 35" in length and coated with halocarbon wax on the interior to decrease

N2O5 losses to the reactor walls. N2O5 was produced in situ via the reaction of excess NO2 with ozone, created via photolysis of O2, in a mixing cell. N2O5 was delivered to the flow reactor via

1/8" OD Teflon tubing, and was injected via a side injection port to either the top or the base of the reactor. N2O5 was monitored as I•N2O5 (m/Q 234) using a quadrupole chemical ionization mass spectrometer (CIMS) running in negative ion mode, with methyl iodide as the reagent ion, as described in Kercher et al., 2009.44

The resulting flow emerging from the base of the flow reactor was split between the

CIMS, sampling 2 slpm, ATOFMS, sampling 1 slpm and a scanning mobility particle sizing instrument (SMPS), sampling 0.4 slpm. During an hour of each sampling day, the N2O5 source was directed to a waste line, and only ambient particles were allowed to pass through the flow reactor. This allowed the ATOFMS to probe unreacted ambient particles.

38

Data obtained during periods of NO ≥ 1ppb was excluded from the gamma calculation, as high NO concentrations can cause titration of NO3, resulting in a shift in the NO3-N2O5 equilibrium, and thus N2O5 signal artifacts.

2.8.2.2 Sizing Measurements:

During the Scripps Pier sampling period, sizing measurements between 0.01-0.6µm were made using a Scanning Mobility Particle Sizer (TSI Inc. 3010 CPC & TSI Inc. 3081 DMA), connected to the ATOFMS sampling line. Particle number distributions measured with the SMPS were then converted into surface area distributions, which were used in the calculation of gamma.

The SMPS measured wet size distributions, so no correction or growth factors were applied to the data.

Sub-micrometer surface area was calculated using the wet size distributions from the

SMPS data. The super-micrometer surface area for both the more and less polluted sampling periods of the Scripps Pier data set was estimated using a similar data set from a separate prior

(August 8th-September 1st) sampling period. During the prior study, both an Aerodynamic

Particle Sizer (APS) and SMPS were measuring size distributions from Scripps Pier. Periods of similar SMPS size distributions were found in the current and prior SMPS data sets. The APS surface area contribution (between 9-24%) for the corresponding time period in the prior study was obtained, and applied to the current data set as an estimate for the super-micrometer particle fraction. Estimation of the super-micrometer surface area reaching the flow reactor may be an overestimation, since the inlet was curved, likely leading to loss of the larger particles due to impact on the inlet wall. For this reason, the reported measured gamma values are likely to be the lower limit.

39

2.8.2.3 Single Particle Composition Measurements:

For all studies reported, aerosol single particle chemical composition was measured using an Aerosol Time of Flight Mass Spectrometer (ATOFMS) which has been described previously.29

Briefly, aerosol particles enter the instrument, where they then encounter a UV laser pulse

(266nm) and are consequently desorbed and ionized. Both positive and negative ions are then detected by a dual time-of-flight mass spectrometer. The ATOFMS detection sensitivity of Cl-

- 45 and NO3 was considered to be approximately equal based on their similar electron affinities.

This is discussed in more detail in the next section. The ATOFMS employed measured particles between 0.1-1.67 m in geometric diameter. Geometric physical diameters were calculated from aerodynamic diameter by assuming particles were spherical, and using the following equation

(ES1):

(ES1)

3 3 where ρeff is the effective density of particles, taken to be 1.8 g/cm , ρ0 is 1g/cm , the density of particles for which the ATOFMS is calibrated. The size range measured by the ATOFMS allows for measurement of the surface area maximum of ~350nm (more polluted), and ~300nm (less polluted).

Two criteria were employed for filtering particles for analysis. Only particles that

- - contained chloride were considered, and from the particles remaining, those with nCl /nNO3

4 - - between 1x10 and 1000 were divided into 500 nCl /nNO3 bins and used to calculate the cumulative distribution function for the data set. The percentage of the original particle population remaining after each filter criteria are shown in Table 2.2. It should be noted that the reason for a large portion of particles being removed before analysis in the nascent SSA case is

40 due to the majority of particles containing large amounts of chloride in comparison to nitrate,

- - causing the nCl /nNO3 to be much greater than 1000.

Table 2.2 The fraction of particles removed before CDF and/or N2O5 calculation.

- - Location Particles containing chloride 0.0001

SIO Wave Channel; nascent SSA 99.8% 26.2%

Jeju Island, Korea 95.6% 89.2%

SIO Pier; polluted 78.5% 76.5%

SIO Pier; less polluted 65.8% 58.1%

2.8.3 Calculations and Interpretations

2.8.3.1 Measurements and Interpretation of Particulate Chloride-to-Nitrate Mixing State:

- - The PDF of particulate nCl /nNO3 for particles sampled at two ambient locations (Jeju

Island, Korea and Long Beach*, California) and under two different laboratory conditions

(generated nascent sea spray, and chemically processed generated nascent sea spray) are shown in

Fig. 2.6. The PDF is employed as a means to examine both the extent of particle aging between populations and the degree of external mixing within a specific particle population. The PDF

- - obtained from particulate nCl /nNO3 measured on individual nascent SSA (blue line, described above) indicates that greater than 50% of particles exhibit a ratio greater than 100, reflecting the

- - nCl /nNO3 of seawater. As a proxy for aging of particles in the atmosphere, nascent SSA were passed through a flow reactor (red line), resulting in a HNO3 exposure of 155 ppb over an average reaction time of 114 s. The resultant PDF associated with HNO3 exposure shows a marked shift

- - towards lower values of nCl /nNO3 as compared with the nascent SSA PDF, illustrating the result of chemical aging on a population of nascent SSA. This decrease in the ensemble average

41

- - particulate nCl /nNO3 would result in a corresponding suppression of N2O5 uptake when calculated using E1. Ambient data, collected over 8 days, from both Jeju Island and Long Beach,

(black and grey curves, respectively), exhibit differing degrees of ambient aging based on the

- - location of the PDF curve with respect to nCl /nNO3 . In comparison to the fresh nascent SSA, the

- - Long Beach particle population is shifted to lower values of nCl /nNO3 , and thus is more aged.

Following the same reasoning, the Jeju Island particle population appears more aged than nascent sea spray, but less aged than the particles measured in Long Beach. This is consistent with the proximity of the sampling location to nascent SSA emissions in the absence of strong anthropogenic influence. In addition to providing information on particle age with respect to

- -, emission (mean value), the observed PDF of nCl /nNO3 when computed on time-scales that are short relative to atmospheric variability, provides an indication of the mixing state of the particle population.

42

Ambeint (Long Beach,CA) HNO3-Reacted SSA -3 12x10 Ambient (Jeju Island,Korea) Nascent SSA

10

8 PDF 6

4

2

0 0.0001 0.001 0.01 0.1 1 10 100 1000 - - nCl /nNO3

- - Figure 2.6 Probability distribution curves of particulate nCl /nNO3 calculated for four cases: ambient aerosol from Jeju Island, Korea (grey) and Long Beach, CA (black), nascent sea spray aerosol generated in the SIO wave channel (blue), and the same nascent sea spray after reaction with HNO3 (red).

2.8.3.2 Calculation of k'2f:

The value of k'2f used in the parameterization to calculate N2O5 at the SIO pier was

1.25×106 s-1 for the more polluted period, and 1.13×106s-1 for the less polluted period. According to the equation given in Bertram et al. 2009,30 this value depends on the liquid water content of the aerosol particles. Relative humidity for sampling periods 1 and 2 of the Scripps Pier study was obtained from a meteorological station at a nearby coastal site due to absent RH data from the Scripps Pier meteorological station during these time periods. Using this data, the Aerosol

Inorganics Model31 was utilized to determine the liquid water content of particles at an average relative humidity of 71% for the more polluted period, and 64% for the less polluted period. k'2f

43 for the wave channel nascent SSA data was calculated in a similar way, also using a relative humidity of 64%.

- Calculation of N2O5 using the extreme values from the nH2O/nNO3 range had a larger impact on the calculated gamma for the more polluted condition, producing at most a 28% error.

A similar calculation for the less polluted condition lead to at most a 10% difference in N2O5.

These values are summarized in Table 2.1 of the main paper.

2.8.3.3 Probability Distribution Function:

The shape of the cumulative distribution function is a useful tool in the assessment of particle population age and the degree to which the sampled population is externally mixed.

However, the width of the PDF is a complex function of: 1) single particle mixing state, 2)

ATOFMS instrument precision, and 3) temporal variability on the time-scale of the observation.

- - As shown in Fig. 2.6, the width of the PDF curves are highly variable, where nCl /nNO3 as

- - measured in Long Beach is narrowly distributed and nCl /nNO3 measured at Jeju Island is broadly distributed between 1x10-2 and 1x102. In what follows, we assess the contribution of

ATOFMS accuracy and precision and temporal variability in the measurements to the width of

- - the nCl /nNO3 distributions, using the Jeju island samples as an example.

- - Figure 2.7A shows the single particle nCl /nNO3 for particles sampled from Jeju Island over a 24 hour period (light grey markers). The large variability in the single particle data can be deconvoluted, to some extent, via the analysis of standard solutions prepared in the lab and analyzed using an ATOFMS (Figure 2.7B). Standard I is aerosol produced via atomization of a

10:80:10 solution of NaCl:NaNO3:NaI in water. Standard II is aerosol produced via atomization of a 45:45:10 solution of NaCl:NaNO3:NaI in water. NaI is added to the water to increase particle absorption of the 266 nm ionization laser pulses. Analysis of these solutions shows a range of

- - nCl /nNO3 , with the distributions tending towards lower ratios than expected based on the ratio of

44 components in solution. This bias is likely due to matrix effects in the ATOFMS ionization

- 45 process, caused by the slightly differing electron affinities between Cl and NO3. To determine the significance of this bias, the interquartile range of the median of single particle data for

Standard II, 65%, was applied as a percentage to the running median of the Jeju Island data (Fig.

2.7A blue line) resulting in the solid dark grey region in Figure 2.7A. This analysis indicates the

ATOFMS bias determined from standard solutions is significantly smaller than the interquartile

- - variability seen in the data set (light solid gray region). Figure 2.7C shows particulate nCl /nNO3 as a function of particle size and frequency of occurrence, with the particulate surface area

- - maximum (Figure 2.7D) deriving from particles with nCl /nNO3 of 0.25. It should be noted that with an ATOFMS size detection limit of 0.1-1.67 m, the percentage of particle surface area

- - captured in this data set is 74%. An increase in particle nCl /nNO3 is seen with increasing size,

- - providing additional evidence of contribution to the variability in the nCl /nNO3 arising from particle-to-particle differences.

45

Figure 2.7 Assessment of variation in the cumulative distribution function arising from sources unrelated - - to particle-to-particle variability. Panel A shows the raw variability in nCl /nNO3 over 24 hours of sampling at Jeju Island (light grey markers). The running median over 500s of sampling is shown in blue. - - Applying the nCl /nNO3 interquartile range, observed in a standard solution, as a percentage to the running - - median results in the solid dark grey area. Panel B shows nCl /nNO3 PDF curves of two standard solutions (Standard I, 10:80:10 of NaCl:NaNO3:NaI in water; Standard II, 40:40:10 of NaCl:NaNO3:NaI in water), - - along with the PDF of ambient Jeju Island data. Panel C: particulate nCl /nNO3 as a function of particle size and frequency of occurrence. Panel D: surface area data for Jeju Island (blue line), with standard deviation shown in blue shading.

2.8.3.4 Internally Mixed (Ensemble Average Assumption) N2O5 Calculation:

The ensemble average N2O5 calculation is derived by handling the chloride and nitrate concentrations of particles as if they had been collected on a filter sample. The sum of total chloride and total nitrate is calculated for the entire particle population from the single particle

46 measurements. An average chloride and an average nitrate value is obtained, and then the average chloride value divided by the average nitrate value in order to determine an ensemble average

- - nCl /nNO3 for the population. This ensemble average value is then used in the N2O5 parameterization to calculate a particle population ensemble average gamma value.

The error for the ensemble average calculation was calculated by propagating the error in

- - the ensemble average nCl /nNO3 calculation. This propagated error was then used to calculate a

- - lower and upper limit to the nCl /nNO3 , and these values used in the parameterization to determine the limits on the N2O5 value.

2.8.3.5 Externally Mixed N2O5 Calculation:

- - The externally mixed N2O5 values are derived from the particle population nCl /nNO3

- - probability distribution function (PDF). For each of the 500 nCl /nNO3 bins used in the PDF calculation, a midpoint is established. This is then used to calculate the N2O5 value of each bin, and is multiplied by the population of particles in that respective bin. The externally mixed N2O5 average value is then calculated by summing over the N2O5 values determined for each bin. The upper and lower error values were calculated utilizing the Standard I and II atomized aerosol. A probability distribution function of each standard was made and fitted with a log normal fit curve.

- - The nCl /nNO3 corresponding to the peak value of the curve was determined, and the percent

- - deviation from the theoretical maximum nCl /nNO3 value ( 0.125 for standard I, and 1 for standard II) found. For standard I, the percent deviation was 68.6%, and was 56.4% for standard

- - - - II. The upper nCl /nNO3 value was considered to vary by +68.6%. Upper nCl /nNO3 values were input into the parameterization, resulting in an upper bound for the externally mixed gamma calculations.

Standards I and II were also used to calculate an additional upper error bound, which was added to the previously calculated value, and a lower bound. The full width half maximum

47

(FWHM) for each standard PDF curve was determined. The x-values corresponding to the

FWHM spread were used to calculate a sigma value ( = 0.5 × FWHM). Sigma was calculated as

- - a percentage of the FWHM nCl /nNO3 value, resulting in 79.6% for Standard I, and 80.6% for

- - standard II. Assuming each nCl /nNO3 bin could differ by as much as 80.6% in the externally

- - mixed gamma calculation, this value was applied to the nCl /nNO3 ratios, and used in the parameterization to calculate a lower, and an additional upper bound. The upper bound error from this calculation was combined with the value previously described.

- 2.8.3.6 Calculation of nH2O/nNO3

- 31 nH2O/nNO3 was determined using the Aerosol Inorganics Model, principal model III.

Inputs used with AIM were the mean RH determined for Periods 1 and 2, and moles of particulate

- - Cl and NO3 with Na as a counter ion. Organics were not used as inputs into the model for a number of reasons: 1) We do not have specific mass fractions of organics on each ambient particle to include in the model input, 2) the chemical identity of the specific organic species on the ambient particles is unknown , 3) the phase state and spatial distribution in/on the particles themselves are not known, and 4) this study was intended to focus on the inorganic mixing state and its influence on the discrepancy between measured and parameterized values of (N2O5). For

- - - - both RH conditions, the model was run for nCl /nNO3 equal to 0.0001 (the smallest nCl /nNO3 used in the parameterization calculation) and 1000 (the maximum used in the parameterization.

- - - From the AIM output, nH2O/nNO3 was determined for each nCl /nNO3 , and a plot of

- - - nH2O/nNO3 vs. nCl /nNO3 created for each of the two RH conditions and fit linearly. The

- equation from the fit for was used in the subsequent calculation of nH2O/nNO3 for each of the

- - - - 500 nCl /nNO3 bins (for externally mixed case) and for the average nCl /nNO3 (for the internally mixed case).

48

2.9 Acknowledgments

Chapter 2, in full, is a reprint of the material as it appears in Environmental Science and

Technology. Olivia S. Ryder, Andrew P. Ault, John F. Cahill, Timothy L. Guasco, Theran P.

Riedel, Luis A. Cuadra-Rodriguez, Cassandra J. Gaston, Elizabeth Fitzgerald, Christopher Lee,

Kimberly A. Prather, and Timothy H. Bertram. (2014), On the Role of Particle Inorganic Mixing

State in the Reactive Uptake of N2O5 to Ambient Aerosol Particles, Environ. Sci. Technol., 2014,

48 (3), pp 1618–1627, doi: 10.1021/es4042622. The dissertation author was the primary investigator and author of this paper.

The authors would like to thank Joel Thornton for the use of the CIMS and flow tube apparatus used in the ambient portion of this study. This research was supported by the Office of

Science (Office of Biological and Environmental Research), U.S. Department of Energy (Grant

No. DE-SC0006431) and the National Science Foundation (Grant No. CHE1038028). O.S.R is grateful for a Graduate Research Fellowship from the National Science Foundation (2011-2014).

2.10 References

(1) Riedel, T. P.; Bertram, T. H.; Ryder, O. S.; Liu, S.; Day, D. A.; Russell, L. M.; Gaston, C. J.; Prather, K. A.; Thornton, J. A. Direct N2O5 Reactivity Measurements at a Polluted Coastal Site. Atmos Chem Phys 2012, 12, 2959–2968.

(2) Riedel, T. P.; Bertram, T. H.; Crisp, T. A.; Williams, E. J.; Lerner, B. M.; Vlasenko, A.; Li, S.-M.; Gilman, J.; de Gouw, J.; Bon, D. M.; Wagner, N. L.; Brown, S. S.; Thornton, J. A. Nitryl Chloride and Molecular Chlorine in the Coastal Marine Boundary Layer. Environ. Sci. Technol. 2012, 46, 10463–10470.

(3) Brown, S. S.; Dubé, W. P.; Fuchs, H.; Ryerson, T. B.; Wollny, A. G.; Brock, C. A.; Bahreini, R.; Middlebrook, A. M.; Neuman, J. A.; Atlas, E.; Roberts, J. M.; Osthoff, H. D.; Trainer, M.; Fehsenfeld, F. C.; Ravishankara, A. R. Reactive Uptake Coefficients for N2O5 Determined from Aircraft Measurements during the Second Texas Air Quality Study: Comparison to Current Model Parameterizations. J. Geophys. Res. Atmospheres 2009, 114, D00F10.

(4) Bertram, T. H.; Thornton, J. A.; Riedel, T. P.; Middlebrook, A. M.; Bahreini, R.; Bates, T. S.; Quinn, P. K.; Coffman, D. J. Direct Observations of N2O5 Reactivity on Ambient Aerosol Particles. Geophys. Res. Lett. 2009, 36, L19803.

49

(5) Brown, S. S.; Ryerson, T. B.; Wollny, A. G.; Brock, C. A.; Peltier, R.; Sullivan, A. P.; Weber, R. J.; Dubé, W. P.; Trainer, M.; Meagher, J. F.; Fehsenfeld, F. C.; Ravishankara, A. R. Variability in Nocturnal Nitrogen Oxide Processing and Its Role in Regional Air Quality. Science 2006, 311, 67–70.

(6) Folkers, M.; Mentel, T. F.; Wahner, A. Influence of an Organic Coating on the Reactivity of Aqueous Aerosols Probed by the Heterogeneous Hydrolysis of N2O5. Geophys. Res. Lett. 2003, 30, 1644.

(7) Anttila, T.; Kiendler-Scharr, A.; Tillmann, R.; Mentel, T. F. On the Reactive Uptake of Gaseous Compounds by Organic-Coated Aqueous Aerosols: Theoretical Analysis and Application to the Heterogeneous Hydrolysis of N2O5. J. Phys. Chem. A 2006, 110, 10435–10443.

(8) McNeill, V. F.; Patterson, J.; Wolfe, G. M.; Thornton, J. A. The Effect of Varying Levels of Surfactant on the Reactive Uptake of N2O5 to Aqueous Aerosol. Atmos Chem Phys 2006, 6, 1635–1644.

(9) Park, S.-C.; Burden, D. K.; Nathanson, G. M. The Inhibition of N2O5 Hydrolysis in Sulfuric Acid by 1-Butanol and 1-Hexanol Surfactant Coatings. J. Phys. Chem. A 2007, 111, 2921–2929.

(10) Knopf, D. A.; Cosman, L. M.; Mousavi, P.; Mokamati, S.; Bertram, A. K. A Novel Flow Reactor for Studying Reactions on Liquid Surfaces Coated by Organic Monolayers: Methods, Validation, and Initial Results. J. Phys. Chem. A 2007, 111, 11021–11032.

(11) Cosman, L. M.; Knopf, D. A.; Bertram, A. K. N2O5 Reactive Uptake on Aqueous Sulfuric Acid Solutions Coated with Branched and Straight-Chain Insoluble Organic Surfactants. J. Phys. Chem. A 2008, 112, 2386–2396.

(12) Abbatt, J. P. D.; Lee, A. K. Y.; Thornton, J. A. Quantifying Trace Gas Uptake to Tropospheric Aerosol: Recent Advances and Remaining Challenges. Chem. Soc. Rev. 2012, 41, 6555–6581.

(13) Thornton, J. A.; Braban, C. F.; Abbatt, J. P. D. N2O5 Hydrolysis on Sub-Micron Organic Aerosols: The Effect of Relative Humidity, Particle Phase, and Particle Size. Phys. Chem. Chem. Phys. 2003, 5, 4593–4603.

(14) Wahner, A.; Mentel, T. F.; Sohn, M.; Stier, J. Heterogeneous Reaction of N2O5 on Sodium Nitrate Aerosol. J. Geophys. Res. Atmospheres 1998, 103, 31103–31112.

(15) Behnke, W.; George, C.; Scheer, V.; Zetzsch, C. Production and Decay of ClNO2 from the Reaction of Gaseous N2O5 with NaCl Solution: Bulk and Aerosol Experiments. J. Geophys. Res. Atmospheres 1997, 102, 3795–3804.

(16) Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd Edition; Wiley-Interscience, 2006.

50

(17) Murphy, D. M.; Cziczo, D. J.; Froyd, K. D.; Hudson, P. K.; Matthew, B. M.; Middlebrook, A. M.; Peltier, R. E.; Sullivan, A.; Thomson, D. S.; Weber, R. J. Single-Particle Mass Spectrometry of Tropospheric Aerosol Particles. J. Geophys. Res. Atmospheres 2006, 111, D23S32.

(18) Zaveri, R. A.; Barnard, J. C.; Easter, R. C.; Riemer, N.; West, M. Particle-Resolved Simulation of Aerosol Size, Composition, Mixing State, and the Associated Optical and Cloud Condensation Nuclei Activation Properties in an Evolving Urban Plume. J. Geophys. Res. Atmospheres 2010, 115, D17210.

(19) Marcolli, C.; Luo, B. P.; Peter, T.; Wienhold, F. G. Internal Mixing of the Organic Aerosol by Gas Phase Diffusion of Semivolatile Organic Compounds. Atmos Chem Phys 2004, 4, 2593–2599.

(20) 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, 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, 1525–1529.

(21) Finlayson-Pitts, B. J.; Jr, J. N. P. Chemistry of the Upper and Lower Atmosphere: Theory, Experiments, and Applications; Academic Press, 1999.

(22) Keene, W. C.; Savoie, D. L. The pH of Deliquesced Sea-Salt Aerosol in Polluted Marine Air. Geophys. Res. Lett. 1998, 25, 2181–2184.

(23) Pratt, K. A.; Hatch, L. E.; Prather, K. A. Seasonal Volatility Dependence of Ambient Particle Phase Amines. Environ. Sci. Technol. 2009, 43, 5276–5281.

(24) Bertram, A. K.; Martin, S. T.; Hanna, S. J.; Smith, M. L.; Bodsworth, A.; Chen, Q.; Kuwata, M.; Liu, A.; You, Y.; Zorn, S. R. Predicting the Relative Humidities of Liquid- Liquid Phase Separation, Efflorescence, and Deliquescence of Mixed Particles of Ammonium Sulfate, Organic Material, and Water Using the Organic-to-Sulfate Mass Ratio of the Particle and the Oxygen-to-Carbon Elemental Ratio of the Organic Component. Atmos Chem Phys 2011, 11, 10995–11006.

(25) Meng, Z.; Seinfeld, J. H. Time Scales to Achieve Atmospheric Gas-Aerosol Equilibrium for Volatile Species. Atmos. Environ. 1996, 30, 2889–2900.

(26) Jacobson, M. Z. A Solution to the Problem of Nonequilibrium Acid/Base Gas-Particle Transfer at Long Time Step. Aerosol Sci. Technol. 2005, 39, 92–103.

51

(27) Liu, Y.; Cain, J. P.; Wang, H.; Laskin, A. Kinetic Study of Heterogeneous Reaction of Deliquesced NaCl Particles with Gaseous HNO 3 Using Particle-on-Substrate Stagnation Flow Reactor Approach. J. Phys. Chem. A 2007, 111, 10026–10043.

(28) Bertram, T. H.; Thornton, J. A.; Riedel, T. P. An Experimental Technique for the Direct Measurement of N2O5 Reactivity on Ambient Particles. Atmos Meas Tech 2009, 2, 231– 242.

(29) Gard, E.; Mayer, J. E.; Morrical, B. D.; Dienes, T.; Fergenson, D. P.; Prather, K. A. Real- Time Analysis of Individual Atmospheric Aerosol Particles: Design and Performance of a Portable ATOFMS. Anal. Chem. 1997, 69, 4083–4091.

(30) Bertram, T. H.; Thornton, J. A. Toward a General Parameterization of N2O5 Reactivity on Aqueous Particles: The Competing Effects of Particle Liquid Water, Nitrate and Chloride. Atmos Chem Phys 2009, 9, 8351–8363.

(31) Clegg, S. L.; Brimblecombe, P.; Wexler, A. S. Thermodynamic Model of the System H + + + 2- - - −NH 4 −Na −SO 4 −NO 3 −Cl −H 2 O at 298.15 K. J. Phys. Chem. A 1998, 102, 2155–2171.

(32) Draxler, R. R; Rolph, G. D. HYSPLIT (HYbrid Single-Particle Lagrangian Integrated Trajectory) Model Access via NOAA ARL READY Website (http://ready.arl.noaa.gov/HYSPLIT.php); NOAA Air Resources Laboratory, Silver Spring, MD, 2013.

(33) Rolph, G. D. Real-Time Environmental Applications and Display sYstem (READY) Website (http://ready.arl.noaa.gov); NOAA Air Resources Laboratory, Silver Spring, MD, 2013. (34) Hawkins, L. N.; Russell, L. M.; Covert, D. S.; Quinn, P. K.; Bates, T. S. Carboxylic Acids, Sulfates, and Organosulfates in Processed Continental Organic Aerosol over the Southeast Pacific Ocean during VOCALS-REx 2008. J. Geophys. Res. Atmospheres 2010, 115, D13201.

(35) Thornton, J. A.; Abbatt, J. P. D. N2O5 Reaction on Submicron Sea Salt Aerosol: Kinetics, Products, and the Effect of Surface Active Organics. J. Phys. Chem. A 2005, 109, 10004– 10012.

(36) Silva, P. J.; Prather, K. A. Interpretation of Mass Spectra from Organic Compounds in Aerosol Time-of-Flight Mass Spectrometry. Anal. Chem. 2000, 72, 3553–3562.

(37) Kuwata, M.; Martin, S. T. Phase of Atmospheric Secondary Organic Material Affects Its Reactivity. Proc. Natl. Acad. Sci. 2012, 109, 17354–17359.

(38) Saukko, E.; Lambe, A. T.; Massoli, P.; Koop, T.; Wright, J. P.; Croasdale, D. R.; Pedernera, D. A.; Onasch, T. B.; Laaksonen, A.; Davidovits, P.; Worsnop, D. R.; Virtanen, A. Humidity-Dependent Phase State of SOA Particles from Biogenic and Anthropogenic Precursors. Atmos Chem Phys 2012, 12, 7517–7529.

52

(39) Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; DeMott, P. J.; Aluwihare, L. I.; Palenik, B. P.; Azam, F.; Seinfeld, J. H.; Moffet, R. C.; Molina, M. J.; Cappa, C. D.; Geiger, F. M.; Roberts, G. C.; Russell, L. M.; Ault, A. P.; Baltrusaitis, J.; Collins, D. B.; Corrigan, C. E.; Cuadra-Rodriguez, L. A.; Ebben, C. J.; Forestieri, S. D.; Guasco, T. L.; Hersey, S. P.; Kim, M. J.; Lambert, W. F.; Modini, R. L.; Mui, W.; Pedler, B. E.; Ruppel, M. J.; Ryder, O. S.; Schoepp, N. G.; Sullivan, R. C.; Zhao, D. Bringing the Ocean into the Laboratory to Probe the Chemical Complexity of Sea Spray Aerosol. Proc. Natl. Acad. Sci. 2013, 110, 7550–7555.

(40) Ault, A. P.; Moffet, R. C.; Baltrusaitis, J.; Collins, D. B.; Ruppel, M. J.; Cuadra- Rodriguez, L. A.; Zhao, D.; Guasco, T. L.; Ebben, C. J.; Geiger, F. M.; Bertram, T. H.; Prather, K. A.; Grassian, V. H. Size-Dependent Changes in Sea Spray Aerosol Composition and Properties with Different Seawater Conditions. Environ. Sci. Technol. 2013, 47, 5603–5612.

(41) Deane, G. B.; Stokes, M. D. Scale Dependence of Bubble Creation Mechanisms in Breaking Waves. Nature 2002, 418, 839–844.

(42) Bauer, S. E.; Ault, A.; Prather, K. A. Evaluation of Aerosol Mixing State Classes in the GISS modelE-MATRIX Climate Model Using Single-Particle Mass Spectrometry Measurements. J. Geophys. Res. Atmospheres 2013, 118, 9834–9844.

(43) Ault, A. P.; Gaston, C. J.; Wang, Y.; Dominguez, G.; Thiemens, M. H.; Prather, K. A. Characterization of the Single Particle Mixing State of Individual Ship Plume Events Measured at the Port of Los Angeles. Environ. Sci. Technol. 2010, 44, 1954–1961.

(44) J. P. Kercher, T. P. R. Chlorine Activation by N2O5: Simultaneous, in Situ Detection of ClNO2 and N2O5 by Chemical Ionization Mass Spectrometry. Atmospheric Meas. Tech. 2009, 2.

(45) Linstrom, P. J.; Mallard, W. G. In NIST Standard Reference Database Number 69 [Online]; Eds. National Institute of Standards and Technology: Gaithersburg MD, 20899.

Chapter 3 On the Role of Organic Surfactants and Coatings in Regulating N2O5 Reactive Uptake to Sea Spray Aerosol

3.1 Abstract

Previous laboratory measurements and field observations have suggested that the reactive uptake of N2O5 to sea spray aerosol particles is a complex function of particle chemical composition and phase, where surface active organics can suppress the reactive uptake by orders of magnitude. However, to date there are no direct studies of the reactive uptake of N2O5 to nascent sea spray aerosol that permit assessment of the role that organic films present at the surface of sea spray aerosol (SSA) may play in suppressing or enhancing N2O5 uptake kinetics. In this study, SSA was generated from ambient seawater and artificial seawater matrices using a

Marine Aerosol Reference Tank (MART), capable of producing nascent SSA representative of ambient conditions. The reactive uptake of N2O5 ((N2O5)) was measured using an entrained aerosol flow reactor coupled to a chemical ionization mass spectrometer for measurement of the surface area (Sa) dependent heterogeneous loss rate (khet). Measurements of (N2O5) for SSA generated from salt water sequentially doped with representative organic mimics, or from ambient seawater do not deviate statistically from that observed for sodium chloride ((N2O5)NaCl

= 0.01-0.03) when relative humidity (RH) is between 49-65 %. The results are consistent with measurements made under clean marine conditions at the Scripps Institution of Oceanography

Pier as well as those conducted on nascent SSA generated in a wave channel. The results presented here suggest that the organic mass fraction of SSA (at RH of 49-65 %) does not limit

(N2O5) as has been suggested previously.

53

54

3.2 Introduction

Heterogeneous reactions play a controlling role in both setting the atmospheric lifetime of aerosol particles and catalyzing the production and loss of trace gases in the atmosphere.1,2 Few heterogeneous reactions have garnered as much interest as the reaction of dinitrogen pentoxide

(N2O5) at aqueous interfaces, as this reaction has been shown to be highly sensitive to surface chemical composition3–5 (see chapter 3), phase,5,6 and the presence of surfactants.7–13 In the atmosphere, accurate representation of the lifetime of N2O5 is of extreme importance as it serves as both a nocturnal reservoir for nitrogen oxides (NOx = NO+ NO2), as well as a pathway for the

14–16 production of photolabile halogen species such as nitryl chloride (ClNO2).

The efficiency of N2O5 reactive uptake at an aqueous interface is commonly reported as the reactive uptake coefficient, (N2O5), or the reaction probability per N2O5 collision with an aerosol particle. The reaction mechanism, as described in the reaction sequence below, begins with the mass accommodation of N2O5 to the aerosol surface. Following accommodation, the

+ autoionization of N2O5 results in the formation of a highly reactive nitronium ion (NO2 )

- intermediate, which reacts promptly with water to form HNO3 (R3), or Cl to form ClNO2 (R4).

N2O5 (g) ↔ N2O5 (aq) (R1)

+ - N2O5 (aq) ↔ NO2 (aq) + NO3 (aq) (R2)

+ + NO2 (aq) + H2O (l) ↔ H (aq) + HNO3 (aq) (R3)

+ - NO2 (aq) + Cl (aq) ↔ ClNO2 (aq) (R4)

ClNO2 (aq) ↔ ClNO2 (g) (R5)

Recent studies have focused on the role the organic mass fraction plays in altering

(N2O5), through either a reduction in the entry probability of N2O5 due to the presence of a

55

7,11 surfactant film, or a diffusive limitation in the N2O5 transport from the interface to the aqueous core of the particle.8 Ambient studies such as Bertram et al. 2009 report a strong anti-correlation between (N2O5) and the particulate organic to sulfate ratio at high relative humidity, consistent with the picture of a phase separated organic film.17,18 Using aircraft observations, Brown et al. observed up to a factor of ten decrease in (N2O5) during ambient measurements in the northeastern United States, correlated with an increase in particulate organic and ammonium sulfate content.19 Laboratory studies using both smog chambers and molecular mimics have demonstrated that organic films exhibiting monolayer coverage can serve to reduce (N2O5) by

8 more than a factor of 60. Escorcia et al. observed a 9-79× reduction in (N2O5) directly dependent on the mass fraction of a-pinene oxidation products in mixed secondary organic

3 aerosol containing ammonium bisulfate. The authors attribute the reduction in (N2O5) observed here to either an organic-induced reduction in mass accommodation coefficient, or an increased hindrance of N2O5 dissolution and diffusion in the particle bulk post-accommodation. Parallel studies have focused on the role of surfactant monolayers in suppressing (N2O5). For example,

Cosman et al. investigated the effect of organic straight chain surfactants of increasing length on the uptake of N2O5 to H2SO4 particles, finding that (N2O5) was suppressed 17-61× as compared

8 to (N2O5) on the sub-phase alone. However, when a branched organic surfactant such as phytanic acid coated the particles, no uptake suppression was observed and they concluded that the results indicate the degree of organic packing on the reactive surface governs the ability of

N2O5 to reach the reactive sub-phase. A 3.5 wt% solution equal to a monolayer coverage of sodium dodecyl sulfate on aqueous NaCl particles was found to effectively suppress the reactive

7 uptake of N2O5 by 10×, while a monolayer coverage of a shorter chain organic, hexanoic acid, on

11 artificial seawater aerosol reduced (N2O5) by 3-4×. Whereas many of these previous studies

56 have focused on secondary organic aerosol, this work concentrates on reactions occurring at the surface of primary sea spray aerosol generated in the laboratory.

In the following manuscript, we focus specifically on the role of organics in nascent SSA in altering (N2O5), as sub-micrometer sized sea spray aerosol have been shown to contain a large mass fraction of organic components (17-83 %).20–22 These organic mass fractions are inconsistent with the early conceptual representation of SSA as salty particulates with well- ordered surfactant films at the surface.7–9 Here, we focus our attention on the role of organic films at the surface of SSA particles in the suppression or enhancement of (N2O5) using a combination of molecular mimics and nascent seawater.

The organic mass fraction of SSA is critically linked to the biological, chemical, and physical processes occurring on the surface ocean.20–22 Specifically the inter-related life cycle of phytoplankton, bacteria, and viruses referred to as the microbial loop,23–26 serves to enhance and transform organic material in surface waters. The primary stages of the microbial loop include an increase in phytoplankton abundance, followed by a subsequent increase in bacterial concentrations commensurate with phytoplankton decay. Bacteria number concentrations eventually decay due to virus grazing, resulting in a spike in virus count. As a result of species cycling, the concentration and speciation of biologically-derived organic material also varies based on the stage of the phytoplankton bloom. It has been suggested that the composition and mass fraction of organics in SSA will track the biological and chemical evolution of a phytoplankton bloom. In this paper, we utilize molecular mimics for organic species, representative of those found in the ocean during a phytoplankton bloom, to assess the impact of

SSA chemical composition on the reactive uptake of N2O5. These results are compared to N2O5 uptake values obtained from SSA generated from natural ambient seawater, particles sampled during clean marine conditions at the Scripps Institution of Oceanography Pier,27,28 as well as nascent SSA generated during a real phytoplankton bloom in a wave channel.

57

3.3 Materials and Methods

3.3.1 N2O5 Generation and Detection

29 N2O5 was generated in situ as described previously. Briefly, ozone, generated by flowing UHP zero air over an exposed UV mercury Penray lamp (Jelight, Irvine, CA), is mixed with NO2 at a stoichiometric ratio of 1:10 to force the N2O5 to NO3 equilibrium to be strongly in favor of N2O5. The combined flows are allowed to mix in a glass mixing cell for approximately two minutes, after which the total flow is 100 sccm and is delivered via 1/8" PFA tubing to the entrained aerosol flow reactor.

The aerosol flow reactor closely resembles that described in Bertram et al., 2009.29 The reactor is a 15cm I.D., 90 cm long halocarbon wax [Halocarbon Products Corporation, Series

600] coated stainless steel tube. Prior to entering the flow reactor, aerosols are conditioned to a relative humidity between 39-65 % by varying the exposure time of the aerosol laden air within a desiccator and measuring the RH at the exit of the desiccator. Similar to Bertram et al., the particle stream was sent either directly to the flow tube or first through a filter assembly equipped with a Whatman supported PTFE membrane filter (Whatman, 5.0 m, TE 38) to remove particles from the air stream. The valve assembly was controlled autonomously via computer controlled pneumatically actuated valves. The particle inlet to the reactor is located orthogonal to the flow tube and N2O5 reagent gas delivery to maximize turbulent mixing between the N2O5 gas and the particle-laden air flow in the entry region. The top of the reactor was equipped with a RH and temperature meter (Vaisala, HMP60 Humidity and Temperature Probe). The bottom cap of the flow tube was equipped with a similar port that allowed for exit of gases and reacted particles to flow through a 1/4" Teflon tube to the inlet of the chemical ionization quadrupole mass spectrometer (CI-QMS). The CI-QMS inlet pull was 1800 sccm, set by a critical orifice in the front end of the instrument. The entirety of this flow came from the exit of the flow reactor.

58

N2O5 and ClNO2 concentrations were monitored continuously using a CI-QMS utilizing

- 30 - I (H2O)n reagent ion chemistry. N2O5 was measured directly as the I (N2O5) adduct at 235 m/z,

- and ClNO2 as the I (ClNO2) adduct at 208 and 210 m/z. Routine measurements of HNO3 were

- also made by measuring the signal intensity at 189 m/z, which corresponds to I (HNO3). Data was collected at 1Hz.

3.3.2 Sea Spray Aerosol Generation and Duty Cycle

Sea spray particles were generated in situ using a Marine Aerosol Reference Tank

(MART) described in detail in Stokes et al. 2013.31 Briefly, a 210 L acrylic tank is filled with 120

L of either artificial seawater, or natural seawater obtained directly from the ocean. The MART is equipped with a recirculating water pump that generates an in tank plunging water sheet. The plunging water sheet permits the formation of subsurface bubble plumes and surface foam patches that accurately mimic those observed in the open ocean from breaking waves. To allow for foam bubble bursting in the tank and to more closely represent wave breaking action in the real ocean, the plunging sheet was modulated in a 10s on, 6s off duty cycle, referred to here as the

"SSA production mode". During SSA production mode, the MART system produces aerosol continuously with constant size and surface area distributions, as shown in Figure 3.1. Given the large headspace volume (90 L) and slow flow rates (3 slpm), the SSA production mode plunging modulation is not resolvable, and thus does not impact the resulting aerosol measurements.

For the experiments described here, the SSA production mode duty cycle was computer controlled, permitting autonomous operation. Following each organic addition and after re-filling the tank with real seawater, the MART lid was secured, and the headspace purged with wet UHP nitrogen, created by flowing N2 through a water bubbler. The system was run with the plunging sheet off for at minimum 60 minutes to ensure full purging of any ambient particles in the MART system. Following this, the SSA production mode was initialized and continued as described above for one hour. During this time, flow from the MART passed through the particle filter prior

59

to reaching the flow reactor to allow the walls to equilibrate with the relative humidity. N2O5 was continuously generated and added to the flow reactor, regardless of the filter state. Fifteen minutes prior to the end of the SSA production mode, the filter is turned off, and the aerosol laden air flow is directed into the flow reactor to equilibrate the reactor walls with particles. At the conclusion of the SSA production mode, plunging is arrested in the MART (Figure 3.2A), after which total particle number concentrations decay to zero as the headspace of the MART is depleted in SSA.31 This part of the cycle is referred to as "SSA decay mode". SSA decay mode persists for 60 minutes, after which time the SSA production mode resumes, and the flow tube flow is returned to the filter state. This duty cycle was repeated three times for each water matrix studied. The complete duty cycle is shown in Figure 3.2 alongside corresponding changes in total particle surface area (red), and sub-micrometer particle surface area (green). As shown in Figure

3.2, the decay in particle concentration within the MART is strongly associated with particle size, with larger particles exhibiting a shorter lifetime than smaller particles within the headspace.31

The size dependent loss rate can be used to isolate sub-micrometer particles from super- micrometer particles for study of the loss of N2O5 to sub-micrometer aerosol as size dependent differences in particle chemical composition are expected.

60

Figure 3.1 A typical surface area distribution for SSA produced in the MART system after applying a growth factor value of 2.3, and adjusting for dilution of the SSA concentration to the sizing instrumentation so as not to saturate the detectors. The black dotted line results from combining UHSAS and APS surface area distributions, and the red dotted line is the data interpolation used over this region.

3.3.3 Molecular Mimics for a Synthetic Phytoplankton Bloom

The experiment was performed in two stages, where first a phytoplankton bloom was synthetically reproduced in the laboratory, and second the MART was filled with ambient ocean water collected from Scripps Institution of Oceanography Pier. The first stage involved filling the tank with an artificial seawater, created by mixing MilliQ water and a sea salt mixture (Sea Salts,

Sigma Aldrich, S9883) to create a matrix with ocean-relevant ion concentrations. This artificial seawater matrix was then sequentially doped with 70 m C of each of the following organic

61 constituents: cholesterol (3β-Hydroxy-5-cholestene, Sigma Aldrich, C8667), galactose (D (+)

Galactose, Sigma Aldrich, G0750), lipopolysaccharides (Lipopolysaccharides from Escherichia coli 0111:B4, Sigma Aldrich, L4130), Albumin protein (Bovine Serum Albumin, Sigma Aldrich,

A2153), and 1,2-Dipalmitoyl-sn-glycero-3-phosphate monosodium salt (DPPA, Avanti Polar

Lipids Inc., #830855P).

Figure 3.2 Plunging duty cycle flag as a function of time (Panel A) where a value of 1 indicates plunging is turned on for SSA production mode, and a value of zero indicates plunging is turned off. Panel B shows the response of the total (red) and sub-micrometer (green) surface area to changes in the plunging duty cycle as a function of time.

62

The sea surface microlayer (SSML) constitutes the top 60-1000 m of the ocean surface32 and is a complex mixture enriched in organic and biological species, including a 15-25× enhancement in bacteria and viruses,33 a 1000× enrichment in dissolved organic matter,33–35 and up to a 50× enrichment in metals including Pb, Cu, Ni, and Fe.36 Following ocean wave breaking, air bubbles produced via air entrainment in the bulk water matrix scavenge organics as they travel up the water column, and produce bubbles and foam atop the SSML. These bubbles eventually burst, ejecting organic material into aerosol particles via jet and film drop formation, with the film breaking process producing the majority of sub-micrometer particles via a mechanism which causes their organic-rich composition.20–22,37 Following Burrows et al., we represent different stages of a phytoplankton bloom with five different classes of biologically-derived organic species.38 Ocean concentrations of total organic carbon are on average 60-70 M C.39 However, the SSML is expected to show a marked organic enrichment, producing much higher organic concentrations that in the average ocean state. For this reason, each organic constituent was added to the water matrix in 70 M concentrations.

The organic chemicals used are listed in order of addition to the water matrix in Table

3.1, along with their molecular weight, carbon to oxygen ratio, and solubility as reported in the literature. The first 70M organic addition was cholesterol, a species that has been previously detected in both the SSML (enriched by up to 15 times) and SSA, and originates primarily from phytoplankton, as noted by Hardy, 1982 and references therein.34 Following this galactose was added, to mimic sugars that have been detected during the growth of bacteria.40 The third addition was Lipopolysaccharides, as these constitute a major component of gram negative bacterial cell membranes41 and are often used as a measure of bacterial concentrations in ocean waters. E.g.42,43

To represent protein and lipids also found in seawater, albumin protein and DPPA were added in

70 M concentrations to the water matrix sequentially.

63

In a second separate experiment, the MART system was purged and filled with Pacific

Ocean water obtained from the end of Scripps Institution of Oceanography Pier (La Jolla, CA).

The water was not doped with any additional constituents, and was run as obtained from the ocean. This real ocean matrix served as stage two of the experiment.

Table 3.1 Synthetic seawater components and their respective properties.

Molecular C:O H:C Water Solubility Weight (g/mol) (mg/mL) Cholesterol 386.66 27:1 46:27 0.002 51 Galactose 180.16 1:1 2:1 684 52 Lipopolysaccharides >3880.02 N/A N/A 553 Albumin 66kDa N/A N/A 4054 DPPA 670.87 4.4:1 1.9:1 insoluble55

3.4 Wave Channel Generated Sea Spray

A series of complementary experiments were conducted in the Scripps Institution of

Oceanography (SIO) 30 m wave channel, where nascent sea spray aerosol were generated from ambient seawater using actual breaking waves. The details concerning the large intensive campaign have been described previously.22 Briefly, a mesocosm bloom was produced in a laboratory environment using real ocean water and within a 30 m wave channel capable of producing waves, which allowed for the production of SSA representative of that from the ocean.

The aim of this effort was to probe the effect of biological activity in ocean waters on sea spray properties. During this campaign, we sampled sea spray into a Pyrex flow reactor through a side arm entrance for improved mixing, and was equipped with a moveable injector for N2O5 delivery into the reactor. The maximum residence time of particles in the flow reactor was two minutes.

N2O5 and ClNO2 were measured using a CI-QMS in the same fashion as described above. The data taken over the course of this experiment has been averaged to provide a single value for

64

(N2O5) when the water used to generate sea spray contained high levels of total organic carbon

(TOC > 283±156 M), and one for when TOC levels were low (TOC < 79±10 M).

3.5 Seawater and Sea Spray Aerosol Characterization

3.5.1 Aerosol Particle Characterization

Aerosol particle number and surface area distributions were measured in real time using an Aerosol Particle Sizer (APS, TSI model 3321), measuring particles from 0.4 to 14 m, and an

Ultra High Sensitivity Aerosol Spectrometer (UHSAS, Droplet Measurement Technologies) sizing particles between 0.06 and 1 m. Particles were first dried to 39-65 % RH using a diffusion drier, and an additional dry nitrogen dilution line was added to prevent saturation of the sizing instrumentation. Dilution factors resulting from nitrogen dilution to sizing instruments were confirmed daily. Additionally, particle growth factors determined via an optical closure procedure

44,45 as previously described were applied to the sizing data for use in the calculation of (N2O5) as the particle stream flowing to the sizing instruments was dried, but the stream which entered the flow reactor was not.

Following the third and final SSA decay mode for each water matrix, size segregated

SSA were generated and collected onto Silicon Nitride (SiN) window substrates housed in a micro-orifice uniform-deposit impactor (MOUDI). Particles were not exposed to driers prior to collection. Two MOUDI stages (stage 5 and 7) were analyzed in detail. Particles collected on stage 5 correspond to wet particle diameters between 1-1.8 m and particles collected on stage 7 correspond to wet particle diameters between 0.33-0.56 m. Here we focus on the stage 7 substrates as the most closely map the surface area maximum for sub-micrometer aerosol analyzed.

65

Figure 3.3 AFM amplitude images, taken at ~20-21 % RH, of a sea spray particle generated in the MART, post-galactose addition showing the whole particle mask (Panel A), and the inorganic core mask (Panel B) used to determine organic volume fractions of imaged particles.

Particles collected using MOUDI stage 7 were imaged using Atomic Force Microscopy

(AFM) to estimate the organic volume fraction of approximately 50 particles for each water matrix sample. AFM imaging was performed with a Molecular Force Probe 3D AFM (Asylum research, Santa Barbara, CA). AC mode images were collected at room RH (20-21 %) with silicon probes (MikroMasch, Model CSC37) with a nominal spring constant of 0.35 N/m and a typical tip radius of curvature of 10 nm. Height, amplitude, and phase images were used to create particle masks over both the core and whole particle and an Asylum research particle analyzer used to determine the volume of the masked areas (Figure 3.3). Organic volume fraction was calculated as the volume of the whole particle minus the core volume, divided by the total particle volume. Particle diameters were measured during AFM analysis and the average measured diameters under 20-21 % RH are listed in Table 3.2. These, rather than wet diameters, were used to determine organic film thickness as the organic volume fractions were also determined for these RH conditions (see below).

66

Table 3.2 Average gamma values obtained during each stage of the synthetic bloom in addition to the range of relative humidities inside the flow reactor over which gamma was measured, and the pH and surface tension of the water matrix. The average organic volume fraction is determined via AFM measurements, and the organic film thickness is derived from the organic volume fraction estimates. See text for details.

3.5.2 Water Characterization

Water samples were taken by skimming the top of the water surface with a clean media bottle following each organic addition. Surface tension was measured by agitating the water

6

8 sample bottles and then transferring collected water into a clean glass dish provided with the

Kruss Tensiometer K11 instrument. Here the Wilhelmy plate method was utilized for the surface tension measurements. Water pH was measured from the collected water samples using a pH meter (IQ Scientific Instruments pH/mV/thermometer #1506471)

3.6 (N2O5) Determinations

Traditional measurements of heterogeneous reaction rates using entrained aerosol flow reactors have focused on measuring the decay in reactant concentration at a fixed surface area concentration by varying the interaction time of the gas with the particles, and calculating the reactive uptake coefficient from the dependence of the heterogeneous loss rate on particle surface area.2,5,8,46,47 Here, we describe a new approach where the gas-particle interaction time is fixed, but the surface area is permitted to decay away in the flow reactor. In this technique, the rise in

N2O5 is directly monitored as a function of the decay in Sa.

67

The MART experiment was designed to include the particle decay mode for two purposes: 1) The decay in Sa provides a unique method for calculating the heterogeneous loss rate autonomously, and 2) The removal rate of SSA within the MART is strongly size dependent, as shown in Stokes et al. 2011, where super-micrometer particles are removed more quickly than sub-micrometer particles. This permits us to time gate our analysis to focus exclusively on sub- micrometer aerosol. Figure 3.4 shows a typical duty cycle during this experiment. Total surface area is constant to within 12 % during SSA production mode (Figure 3.4A and 3.4B), until the plunging is turned off (time = 0s) at which point the surface area begins to decay. By 500s, the surface area contribution from super-micrometer particles is less than 1 % (Figure 3.4A, green trace), while the sub-micrometer surface area accounts for the remaining surface area (Figure

3.4B). For the analysis described below, we constrain our sampling window to the period where the rate of change in surface area is slower than the residence time of the flow reactor (10.3 min).

The full decays were segmented into sections equivalent to the average residence time of the flow reactor (10.3 min). In this analysis we only consider periods where the surface area does not vary by more than 20 % between consecutive periods.

68

Figure 3.4 Panel A shows a representative time trace of total particle surface area (grey) and super- micrometer surface area (green) during a typical experimental cycle, where time=0 corresponds to plunging turned off, thus switching from SSA production mode to SSA decay mode. Panel B shows the corresponding surface area trace for sub-micrometer particles.

During the SSA decay mode, the N2O5 signal increases due to the decreasing surface area available for reaction. Due to the delay in signal response resulting from the residence time in the flow tube, the particle surface area data and N2O5 signal were adjusted by 10.3 minutes to match the beginning of the particle decay mode. We first average the N2O5 signal to 20 seconds to

69

match the APS sizing data time scale. Each N2O5 average point during the particle decay period was then divided by the maximum N2O5 signal in the decay period (corresponding to the point of

2 -3 Sa = 0 m cm ). Lastly, the instantaneous observed loss rate (kobs) was calculated using Equation

1:

E1

where res.time is the residence time in the flow reactor (615 seconds), (N2O5)t are the individual

N2O5 time averaged points during particle decay in the reactor, and (N2O5) highest is the highest

N2O5 value during the period.

The reactive uptake coefficient, (N2O5), was calculated as the slope of a plot of instantaneous kobs (Equation 2) as a function of total surface area (Sa).

E2

where is the mean speed of N2O5 (241.7 m/s), and kwall is the loss rate of N2O5 to the flow reactor wall. This method of (N2O5) determination is illustrated in Figure 3.5, where the circular markers are data points and the grey line is a linear fit to the data providing the slope, which is indicated as (N2O5) in the legend. The y-axis intercept is kwall. Additionally shown are blue lines indicating specific gamma values, for reference.

70

Figure 3.5 Representative plot of instantaneous kobs versus the total surface area in the flow reactor. Data was filtered such that over the length of one period of residence time in the flow reactor, the surface area was not changing by more than 20 % (see text for details).

3.7 Results and Discussion

3.7.1 Aerosol Organic Volume Fraction

SSA particles were analyzed by atomic force microscopy (AFM) to determine the organic volume fraction for particle populations collected from each water matrix. The frequency of occurrence of each organic volume fraction measured per water matrix is shown in Figure 3.6.

Following the addition of 70 M C cholesterol, the mean organic volume fraction was

0.10±0.05. Upon the addition of 70 M C galactose, the mean organic volume fraction increases to 0.18±0.06. For all subsequent organic additions, where the total [C] of the water matrix increases to 350 M, the mean values fall within the narrow range of 0.19-0.21.

71

Figure 3.6 AFM derived organic volume fraction of 0.33-0.56 m particles (measured as 0.21 m at 21 % RH by AFM) collected using a MOUDI, imaged for each water matrix vs. the fraction of occurrence.

Figure 3.6F indicates that the oceanic particle samples show a narrower organic volume fraction distribution with a standard deviation of 0.03, yet with a very similar average value (0.18) to samples post-galactose addition, which range from 0.18-0.21. These values are listed in Table

3.2. Within error, the organic volume fraction values for the synthetic bloom water matrix (post- cholesterol addition) and the real ocean sample are consistent with values stated previously in the literature for non hygroscopic organic volume fractions for similar sized aerosol derived from real ocean water samples (20-29 % organic volume fraction for particles up to 246 nm, wet diameter).48

Using the AFM-derived diameter (dAFM, at 20-21 % RH), the total volume of the particle can be estimated (Vtot). Knowing the organic volume fraction of the particle permits calculation of the volume of just the organic coating (Vcoat), assuming a total phase separation. The difference between these volumes is the volume of the particle core (Vcore = Vtot – Vcoat), from which the

72

1/3 diameter of the particle core can be determined (dcore = 2 × ((0.75π × Vcore) ). Finally, calculated dcore and dAFM can be used to determine the thickness of the organic coating on the particle (thcoat =

(dAFM - dcore)/2). This calculation was performed for each particle type, and the average diameter determined by AFM and calculated film thicknesses are shown in Table 3.2. Within the assumptions made for this calculation, including extrapolation of the measured AFM diameter to a spherical particle, ignoring any spreading of the particle that may have resulted from impaction onto the surface, and the assumption of phase separation of the organics and core, the film thickness determined is remarkably consistent across all samples and is around 8 nm.

Additionally, the organic mass fraction of the particles was estimated from the organic volume fraction determination. For this calculation, it was assumed the particle core had the same density as NaCl (2.2×10-12 g/m3), and due to the unknown phase of the organic fraction, a density between that of graphite and oil (1.6×10-12 g/m3) was used .49,50

3.7.2 Organic Addition Effect on (N2O5)

For both the synthetic bloom and ambient seawater systems, a minimum of three to a maximum of eight determinations of (N2O5) were made for each water matrix. These values were averaged and the resulting (N2O5) is reported in Table 3.2, along with the range of gamma values measured during each water matrix, the propagated error estimation associated with the average (N2O5) value, the relative humidity range measured in the flow reactor, and the measured pH and surface tension from water samples collected from each water matrix.

73

Figure 3.7 (N2O5) as a function of water side analysis metrics including carbon concentration added to the water matrix (A), water pH (B), surface tension (C). All points are colored by the amount of carbon present in the water matrix, and the real ocean sample is colored green.

To assess the dependence of (N2O5) on the composition of the water matrix from which

SSA is being generated, as well as on the organic content of aerosol particles, (N2O5) was plotted versus the concentration of carbon in the water matrix (Figure 3.7A). Here, both the x-axis and color bar are representative of the concentration of carbon added to the water matrix and the color scale is consistent throughout the rest of the panels in Figure 3.7. The error bars indicate the range in (N2O5). Although (N2O5) appears to increase up to [C]=210 M, after the addition of LPS,

74

and then decrease slightly upon further carbon addition, (N2O5) does not show significant variation with the addition of carbon within the range of gamma values determined for each matrix. However, it is possible that N2O5 may undergo a reaction with LPS at the surface, although a mechanism for this is currently unclear. It should also be noted that the 70 M addition of cholesterol results in an average (N2O5) similar to that of the real ocean water sample.

Bulk water pH was measured throughout the experiments and the (N2O5) results as a function of water pH are plotted in Figure 3.7B, where the color bar correlates to the carbon concentration in the water matrix. The water side pH for the synthetic experiment changes from

7.74 for sea salts in MilliQ water, to 8.15 after the addition of all organics. The pH of the ocean water was 7.87, similar to the pH values after the addition of cholesterol, galactose, and LPS (see

Table 3.2). Despite the range of pH values, (N2O5) is invariable with water pH. Surface water samples exhibit wide ranges in surface tension between 52.6 - 69 mN/m, consistent with the addition of organics. However, this has little impact on (N2O5) as shown in Figure 3.7C. It should be noted that Sigma Aldrich sea salt is a seawater evaporate, and thus inherently contains organic material, which explains the lower surface tension value for the sea salt matrix. A possible reason that the data point associated with a [C] = 350 M results in an increase in surface tension relative to the point where [C] = 280 M is a combination of adding albumin to the MART, operating the MART in SSA production mode for an extended sampling period (6 days) due to an instrumental issue, and then adding DPPA. After the initial addition of albumin, the water sample was acquired for surface tension analysis. Following the addition of albumin to the MART, large amounts of foam were generated and over the course of the extended sampling period this foam dissipated, suggesting that organics had been scavenged. Since the albumin water sample was taken prior to running the tank for six days, the surface tension value is low,

75 consistent with an increase in organic content. However, after 6 days, and a dissipation in foam content likely resulting from lower organic water content due to scavenging, DPPA was added to a water matrix likely containing less than the expected 280 M carbon, and thus the surface tension measurement reflects this decrease in organic content with a higher surface tension value.

Interestingly, despite changes in both the organic carbon content and surface tension of the water, the organic volume fraction of particles sampled did not change significantly, though it should be noted that particles collected for the [C]=280 M stage were collected after 6 days of running in

SSA production mode, which may have resulted in lower aerosol organic content. Since reactive uptake was also performed after 6 days with Albumin in the tank, it is not unexpected that

(N2O5) remains relatively constant over the course of the synthetic bloom.

The (N2O5) values obtained here under all organic conditions were invariant and the trend agrees with that observed by Cosman et al., who showed that coating particles with branched chain organics resulted in a (N2O5) that was invariant to that measured on salt particles, presumably due to poor packing of organics on the particle surface allowing for gaps where N2O5 can reach the aqueous core.8 Similarly, the organic molecules used in this study are unlikely to pack closely in well ordered form on the surface of particles due to their complex structures.

Additionally, the (N2O5) values observed in this study agree with those measured during a recent ambient coastal study for particles with similar organic mass fractions, and within the error bars of the real mesocosm this work agrees with (N2O5) estimates made on a real mesocosm system (Figure 3.8).27,28 It is apparent that from the complex nature and abundance of organics present in ocean sea spray, they are unlikely to suppress the reactive uptake of N2O5 on real SSA levels observed in laboratory studies of highly organic-enriched aerosol (Figure 3.8).

76

Figure 3.8 (N2O5) as a function of the organic mass fraction from this work (green, color coded by TOC water content) as compared to values obtained from a real mesocosm bloom generated in the wave channel facility at Scripps Institution of Oceanography in 2011 (yellow), and literature values for purely organic particles (orange)56 and ambient data taken from a coastal environment (blue).27,28

3.8 Conclusions

These results show that (N2O5) does not vary during a simulated phytoplankton bloom and stays consistent, within error, to the uptake measured on MART-generated ocean water aerosol and values measured under ambient coastal conditions, where particles exhibited a similar organic volume fraction. Despite the additions of carbon up to 350 M, the organic volume fraction as measured using AFM remained between 18-21 % after the second addition of organic, and (N2O5) remained between 0.013-0.031 across all additions. These results are consistent with poor packing of organics on the particles surface, allowing for N2O5 access to the aqueous core for reaction. These results indicate that reduced (N2O5) values measured within coastal marine

77 environments are likely due to post-production particle processing in the atmosphere, rather than marine organic content inhibiting N2O5 reactivity.

3.9 Acknowledgements

Chapter 3, in part, is currently being prepared for submission of the material.

Olivia S. Ryder, Nicole R. Campbell, Holly Morris, Matthew J. Ruppel, Alexei Tivanski, and

Timothy H. Bertram. The dissertation author was the primary investigator and author of this material.

This research was supported by the National Science Foundation via the Center for

Aerosol Impact on Climate and the Environment, a Center for Chemical Innovation (NSF

CHE1305427). The authors gratefully acknowledge Dr. Chris Cappa for use of SMPS and growth factor data used in this study and for helpful discussions throughout the experiments. The authors also thank Steven Schill and Sara Forestieri for participating in the collaborative effort of running these experiments. O.S.R gratefully acknowledges a Graduate Research Fellowship from the

National Science Foundation (2011-2014).

3.10 References

(1) Morris, E. D.; Niki, H. Reaction of Dinitrogen Pentoxide with Water. J. Phys. Chem. 1973, 77 (16), 1929–1932.

(2) Mozurkewich, M.; Calvert, J. G. Reaction Probability of N2O5 on Aqueous Aerosols. J. Geophys. Res. Atmospheres 1988, 93 (D12), 15889–15896.

(3) Escorcia, E. N.; Sjostedt, S. J.; Abbatt, J. P. D. Kinetics of N2O5 Hydrolysis on Secondary Organic Aerosol and Mixed Ammonium Bisulfate−Secondary Organic Aerosol Particles. J. Phys. Chem. A 2010, 114 (50), 13113–13121.

(4) Gross, S.; Iannone, R.; Xiao, S.; Bertram, A. K. Reactive Uptake Studies of NO3 and N2O5 on Alkenoic Acid, Alkanoate, and Polyalcohol Substrates to Probe Nighttime Aerosol Chemistry. Phys. Chem. Chem. Phys. PCCP 2009, 11 (36), 7792–7803. (5) Griffiths, P. T.; Badger, C. L.; Cox, R. A.; Folkers, M.; Henk, H. H.; Mentel, T. F. Reactive Uptake of N2O5 by Aerosols Containing Dicarboxylic Acids. Effect of Particle Phase, Composition, and Nitrate Content. J. Phys. Chem. A 2009, 113 (17), 5082–5090.

78

(6) Thornton, J. A.; Braban, C. F.; Abbatt, J. P. D. N2O5 Hydrolysis on Sub-Micron Organic Aerosols: The Effect of Relative Humidity, Particle Phase, and Particle Size. Phys. Chem. Chem. Phys. 2003, 5 (20), 4593–4603.

(7) McNeill, V. F.; Patterson, J.; Wolfe, G. M.; Thornton, J. A. The Effect of Varying Levels of Surfactant on the Reactive Uptake of N2O5 to Aqueous Aerosol. Atmos Chem Phys 2006, 6 (6), 1635–1644.

(8) Cosman, L. M.; Knopf, D. A.; Bertram, A. K. N2O5 Reactive Uptake on Aqueous Sulfuric Acid Solutions Coated with Branched and Straight-Chain Insoluble Organic Surfactants. J. Phys. Chem. A 2008, 112 (11), 2386–2396.

(9) Anttila, T.; Kiendler-Scharr, A.; Tillmann, R.; Mentel, T. F. On the Reactive Uptake of Gaseous Compounds by Organic-Coated Aqueous Aerosols: Theoretical Analysis and Application to the Heterogeneous Hydrolysis of N2O5. J. Phys. Chem. A 2006, 110 (35), 10435–10443.

(10) Folkers, M.; Mentel, T. F.; Wahner, A. Influence of an Organic Coating on the Reactivity of Aqueous Aerosols Probed by the Heterogeneous Hydrolysis of N2O5. Geophys. Res. Lett. 2003, 30 (12), 1644.

(11) Thornton, J. A.; Abbatt, J. P. D. N2O5 Reaction on Submicron Sea Salt Aerosol: Kinetics, Products, and the Effect of Surface Active Organics. J. Phys. Chem. A 2005, 109 (44), 10004–10012.

(12) Knopf, D. A.; Cosman, L. M.; Mousavi, P.; Mokamati, S.; Bertram, A. K. A Novel Flow Reactor for Studying Reactions on Liquid Surfaces Coated by Organic Monolayers: Methods, Validation, and Initial Results. J. Phys. Chem. A 2007, 111 (43), 11021–11032.

(13) Park, S.-C.; Burden, D. K.; Nathanson, G. M. The Inhibition of N2O5 Hydrolysis in Sulfuric Acid by 1-Butanol and 1-Hexanol Surfactant Coatings. J. Phys. Chem. A 2007, 111 (15), 2921–2929.

(14) Finlayson-Pitts, B. J.; Ezell, M. J.; Pitts, J. N. Formation of Chemically Active Chlorine Compounds by Reactions of Atmospheric NaCl Particles with Gaseous N2O5 and ClONO2. Nature 1989, 337 (6204), 241–244.

(15) Osthoff, H. D.; Roberts, J. M.; Ravishankara, A. R.; Williams, E. J.; Lerner, B. M.; Sommariva, R.; Bates, T. S.; Coffman, D.; Quinn, P. K.; Dibb, J. E.; Stark, H.; Burkholder, J. B.; Talukdar, R. K.; Meagher, J.; Fehsenfeld, F. C.; Brown, S. S. High Levels of Nitryl Chloride in the Polluted Subtropical Marine Boundary Layer. Nat. Geosci. 2008, 1 (5), 324–328. (16) Roberts, J. M.; Osthoff, H. D.; Brown, S. S.; Ravishankara, A. R.; Coffman, D.; Quinn, P.; Bates, T. Laboratory Studies of Products of N2O5 Uptake on Cl− Containing Substrates. Geophys. Res. Lett. 2009, 36 (20), L20808.

79

(17) Bertram, T. H.; Thornton, J. A.; Riedel, T. P.; Middlebrook, A. M.; Bahreini, R.; Bates, T. S.; Quinn, P. K.; Coffman, D. J. Direct Observations of N2O5 Reactivity on Ambient Aerosol Particles. Geophys. Res. Lett. 2009, 36 (19), L19803.

(18) Gaston, C. J.; Thornton, J. A.; Ng, N. L. Reactive Uptake of N2O5 to Internally Mixed Inorganic and Organic Particles: The Role of Organic Carbon Oxidation State and Inferred Organic Phase Separations. Atmos Chem Phys 2014, 14 (11), 5693–5707.

(19) Brown, S. S.; Ryerson, T. B.; Wollny, A. G.; Brock, C. A.; Peltier, R.; Sullivan, A. P.; Weber, R. J.; Dubé, W. P.; Trainer, M.; Meagher, J. F.; Fehsenfeld, F. C.; Ravishankara, A. R. Variability in Nocturnal Nitrogen Oxide Processing and Its Role in Regional Air Quality. Science 2006, 311 (5757), 67–70.

(20) Facchini, M. C.; Rinaldi, M.; Decesari, S.; Carbone, C.; Finessi, E.; Mircea, M.; Fuzzi, S.; Ceburnis, D.; Flanagan, R.; Nilsson, E. D.; de Leeuw, G.; Martino, M.; Woeltjen, J.; O’Dowd, C. D. Primary Submicron Marine Aerosol Dominated by Insoluble Organic Colloids and Aggregates. Geophys. Res. Lett. 2008, 35 (17), L17814.

(21) O’Dowd, C. D.; Facchini, M. C.; Cavalli, F.; Ceburnis, D.; Mircea, M.; Decesari, S.; Fuzzi, S.; Yoon, Y. J.; Putaud, J.-P. Biogenically Driven Organic Contribution to Marine Aerosol. Nature 2004, 431 (7009), 676–680.

(22) Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; DeMott, P. J.; Aluwihare, L. I.; Palenik, B. P.; Azam, F.; Seinfeld, J. H.; Moffet, R. C.; Molina, M. J.; Cappa, C. D.; Geiger, F. M.; Roberts, G. C.; Russell, L. M.; Ault, A. P.; Baltrusaitis, J.; Collins, D. B.; Corrigan, C. E.; Cuadra-Rodriguez, L. A.; Ebben, C. J.; Forestieri, S. D.; Guasco, T. L.; Hersey, S. P.; Kim, M. J.; Lambert, W. F.; Modini, R. L.; Mui, W.; Pedler, B. E.; Ruppel, M. J.; Ryder, O. S.; Schoepp, N. G.; Sullivan, R. C.; Zhao, D. Bringing the Ocean into the Laboratory to Probe the Chemical Complexity of Sea Spray Aerosol. Proc. Natl. Acad. Sci. 2013, 110 (19), 7550–7555.

(23) Russell, L. M.; Hawkins, L. N.; Frossard, A. A.; Quinn, P. K.; Bates, T. S. Carbohydrate- like Composition of Submicron Atmospheric Particles and Their Production from Ocean Bubble Bursting. Proc. Natl. Acad. Sci. 2010, 107 (15), 6652–6657.

(24) Quinn, P. K.; Bates, T. S.; Schulz, K. S.; Coffman, D. J.; Frossard, A. A.; Russell, L. M.; Keene, W. C.; Kieber, D. J. Contribution of Sea Surface Carbon Pool to Organic Matter Enrichment in Sea Spray Aerosol. Nat. Geosci. 2014, 7 (3), 228–232.

(25) Pomeroy, L.; leB. Williams, P.; Azam, F.; Hobbie, J. The Microbial Loop. Oceanography 2007, 20 (2), 28–33.

(26) Azam, F. Microbial Control of Oceanic Carbon Flux: The Plot Thickens. Science 1998, 280 (5364), 694–696. (27) Riedel, T. P.; Bertram, T. H.; Ryder, O. S.; Liu, S.; Day, D. A.; Russell, L. M.; Gaston, C. J.; Prather, K. A.; Thornton, J. A. Direct N2O5 Reactivity Measurements at a Polluted Coastal Site. Atmos Chem Phys 2012, 12 (6), 2959–2968.

80

(28) Ryder, O. S.; Ault, A. P.; Cahill, J. F.; Guasco, T. L.; Riedel, T. P.; Cuadra-Rodriguez, L. A.; Gaston, C. J.; Fitzgerald, E.; Lee, C.; Prather, K. A.; Bertram, T. H. On the Role of Particle Inorganic Mixing State in the Reactive Uptake of N2O5 to Ambient Aerosol Particles. Environ. Sci. Technol. 2014, 48 (3), 1618–1627.

(29) Bertram, T. H.; Thornton, J. A.; Riedel, T. P. An Experimental Technique for the Direct Measurement of N2O5 Reactivity on Ambient Particles. Atmos Meas Tech 2009, 2 (1), 231–242.

(30) Kercher, J. P.; Riedel, T. P.; Thornton, J. A. Chlorine Activation by N2O5: Simultaneous, in Situ Detection of ClNO2 and N2O5 by Chemical Ionization Mass Spectrometry. Atmos Meas Tech 2009, 2 (1), 193–204.

(31) Stokes, M. D.; Deane, G. B.; Prather, K.; Bertram, T. H.; Ruppel, M. J.; Ryder, O. S.; Brady, J. M.; Zhao, D. A Marine Aerosol Reference Tank System as a Breaking Wave Analogue for the Production of Foam and Sea-Spray Aerosols. Atmos Meas Tech 2013, 6 (4), 1085–1094.

(32) Wurl, O.; Obbard, J. P. A Review of Pollutants in the Sea-Surface Microlayer (SML): A Unique Habitat for Marine Organisms. Mar. Pollut. Bull. 2004, 48 (11–12), 1016–1030.

(33) Aller, J. Y.; Kuznetsova, M. R.; Jahns, C. J.; Kemp, P. F. The Sea Surface Microlayer as a Source of Viral and Bacterial Enrichment in Marine Aerosols. J. Aerosol Sci. 2005, 36 (5– 6), 801–812.

(34) Hardy, J. T. The Sea Surface Microlayer: Biology, Chemistry and Anthropogenic Enrichment. Prog. Oceanogr. 1982, 11 (4), 307–328.

(35) Liss, P. S.; Duce, R. A. The Sea Surface and Global Change; Cambridge University Press, 2005.

(36) Piotrowicz, S. R.; Ray, B. J.; Hoffman, G. L.; Duce, R. A. Trace Metal Enrichment in the Sea-Surface Microlayer. J. Geophys. Res. 1972, 77 (27), 5243–5254.

(37) Blanchard, D. C. The Ejection of Drops from the Sea and Their Enrichment with Bacteria and Other Materials: A Review. Estuaries 1989, 12 (3), 127–137.

(38) Burrows, S. M.; Ogunro, O.; Frossard, A. A.; Russell, L. M.; Rasch, P. J.; Elliott, S. M. A Physically Based Framework for Modeling the Organic Fractionation of Sea Spray Aerosol from Bubble Film Langmuir Equilibria. Atmos Chem Phys 2014, 14 (24), 13601–13629.

(39) Hansell, D.; Carlson, C.; Repeta, D.; Schlitzer, R. Dissolved Organic Matter in the Ocean: A Controversy Stimulates New Insights. Oceanography 2009, 22 (4), 202–211. (40) Benner, R.; Herndl, G. Bacterially Derived Dissolved Organic Matter in the Microbial Carbon Pump. In Microbial Carbon Pump in the Ocean, N. Jiao, F. Azam, S. Sanders Eds.; Science/AAAS: Washington, DC, 2011; p 46.

81

(41) Seltmann, G.; Holst, O. The Bacterial Cell Wall; Springer Science & Business Media, 2002.

(42) Nagata, T.; Fukuda, R.; Koike, I.; Kogure, K.; Kirchman, D. L. DEGRADATION BY BACTERIA OF MEMBRANE AND SOLUBLE PROTEIN IN SEAWATER. Aquat. Microb. Ecol. 1998, 14 (1), 29–37.

(43) Maeda, M.; Lee, W. J.; Taga, N. Distribution of Lipopolysaccharide, an Indicator of Bacterial Biomass, in Subtropical Areas of the Sea. Mar. Biol. 1983, 76 (3), 257–262.

(44) Paumgartner, G. Bile Acid Biology and Its Therapeutic Implications: Proceedings of the Falk Symposium 141 (XVIII Internationale Bile Acid Meeting) Held in Stockholm, Sweden, June 18 - 19, 2004; Springer Science & Business Media, 2005.

(45) Gould, S. P. The Final Solubility of D-Galactose in Water. J. Dairy Sci. 1940, 23 (3), 227.

(46) LPS-EB Standard lipopolysaccharide from E. coli 0111:B4 strain- TLR4 ligand http://www.invivogen.com/PDF/LPS_EB_TDS.pdf.

(47) ALBUMIN, BOVINE https://www.sigmaaldrich.com/content/dam/sigma- aldrich/docs/Sigma/Product_Information_Sheet/a4919pis.pdf.

(48) 16:0 PC (DPPC) 1,2-dipalmitoyl-sn-glycero-3-phosphocholine 850355 https://www.avantilipids.com/index.php?option=com_content&view=article&id=216&Ite mid=206&catnumber=850355.

(49) Zhang, X.; Massoli, P.; Quinn, P. K.; Bates, T. S.; Cappa, C. D. Hygroscopic Growth of Submicron and Supermicron Aerosols in the Marine Boundary Layer. J. Geophys. Res. Atmospheres 2014, 119 (13), 2013JD021213.

(50) Atkinson, D. B.; Radney, J. G.; Lum, J.; Kolesar, K. R.; Cziczo, D. J.; Pekour, M. S.; Zhang, Q.; Setyan, A.; Zelenyuk, A.; Cappa, C. D. Aerosol Optical Hygroscopicity Measurements during the 2010 CARES Campaign. Atmos Chem Phys 2015, 15 (8), 4045– 4061.

(51) Stewart, D. J.; Griffiths, P. T.; Cox, R. A. Reactive Uptake Coefficients for Heterogeneous Reaction of N2O5 with Submicron Aerosols of NaCl and Natural Sea Salt. Atmos Chem Phys 2004, 4 (5), 1381–1388.

(52) Fried, A.; Henry, B. E.; Calvert, J. G.; Mozurkewich, M. The Reaction Probability of N2O5 with Sulfuric Acid Aerosols at Stratospheric Temperatures and Compositions. J. Geophys. Res. Atmospheres 1994, 99 (D2), 3517–3532. (53) Park, J. Y.; Lim, S.; Park, K. Mixing State of Submicrometer Sea Spray Particles Enriched by Insoluble Species in Bubble-Bursting Experiments. J. Atmospheric Ocean. Technol. 2013, 31 (1), 93–104.

(54) Royal Society of Chemistry. The Merck Index Online https://www.rsc.org/merck-index (accessed Nov 15, 2014).

82

(55) Haynes, W. M. CRC Handbook of Chemistry and Physics, 95th Edition; CRC Press, 2014.

(56) Badger, C. L.; Griffiths, P. T.; George, I.; Abbatt, J. P. D.; Cox, R. A. Reactive Uptake of N2O5 by Aerosol Particles Containing Mixtures of Humic Acid and Ammonium Sulfate†. J. Phys. Chem. A 2006, 110 (21), 6986–6994.

Chapter 4 On the Role of Organics in Regulating ClNO2 Production at the Air-Sea Interface

4.1 Abstract

We report measurements of the product yield for nitryl chloride (ClNO2) production following the reactive uptake of dinitrogen pentoxide (N2O5) to a wide variety of ambient seawater samples as well as seawater mimics. The ClNO2 yield, as measured for ambient seawater collected from both coastal and open ocean waters, was found to be both insensitive to chlorophyll-a, a marker for biological activity, and significantly lower (0.16-0.30) than that expected for equivalent salt-containing solutions (0.82 ± 0.05). Suppression in the ClNO2 yield can be induced by the addition of aromatic organic compounds (e.g., phenol and humic acid) to synthetic seawater matrices. In the case of phenol, surface tension measurements reveal that the surface phenol:chloride ratio can be enhanced by more than a factor of 103 as compared to bulk ratios for subtle changes in surface tension (< 2 %), providing a mechanism to suppress ClNO2 production at low bulk organic concentrations. These results suggest that ClNO2 production at the air-sea interface may not result in the yield expected for NaCl, and that the reactive uptake of

N2O5 and the subsequent product yield of ClNO2 may serve as a unique probe for the composition of the sea surface microlayer.

4.2 Introduction

Heterogeneous and multiphase reactions play an important role in regulating oxidant loadings in the troposphere, with implications for climate and air quality.1 The reactive uptake of gas-phase N2O5 to atmospheric interfaces has been shown to be an efficient chain termination

2,3 process for nitrogen oxides (NOx ≡ NO2 + NO). Laboratory studies, supported by field

83

84

observations, have shown that the reactive uptake of N2O5 to chloride-containing interfaces can lead to the efficient production of nitryl chloride (ClNO2). Prompt photolysis of ClNO2, following volatilization from the interface, provides not only an efficient channel for chlorine atom production, but also regeneration of NOx.

Laboratory studies have demonstrated efficient conversion of N2O5 to ClNO2 on chloride- containing interfaces representative of those found in the atmosphere ([Cl-] < 1 M), where the

- 4–6 ClNO2 yield ((ClNO2)) reaches 1 for [Cl ] greater than 4.6 M. Efficient ClNO2 production is consistent with a concerted reaction mechanism that is initiated by the prompt hydrolysis of N2O5

+ at the interface (R2), resulting in the production of highly reactive nitronium ions (NO2 ). ClNO2

+ - is thought to be formed following the reaction of NO2 with Cl (R4) that outpaces the reaction of

+ NO2 with H2O (R3).

(R1) N2O5 (g) ↔ N2O5 (aq)

+ - (R2) N2O5 (aq) ↔ NO2 (aq) + NO3 (aq)

+ + (R3) NO2 (aq) + H2O (l) ↔ H (aq) + HNO3 (aq)

+ - (R4) NO2 (aq) + Cl ↔ ClNO2 (aq)

+ (R5) NO2 (aq) + organic (aq) ↔ organic-nitrate (aq)

(R6) ClNO2 (aq) ↔ ClNO2 (g)

- + The competition between Cl and H2O for NO2 has been studied previously and has been

4–6 - parameterized as a function of chloride ion concentration. Beyond H2O and Cl , nitronium ions have been shown to react with aromatic compounds at near the encounter rate.7 This electrophillic substitution reaction has been investigated thoroughly with respect to the nitration of aromatic compounds in solution, e.g. Schofield, 1980, Hoggett et al., 1971 and references within.7,8 In this

+ 2 mechanism, NO2 attacks the sp hybridized carbon of the aromatic ring, forming a carbocation

85 intermediate thus temporarily destroying the aromaticity. This is followed by restoration of aromaticity via loss of H+ from the sp3 hybridized carbon.9

Previous work by Heal et al. demonstrated that the reaction of N2O5 and ClNO2 with

+ 10 phenol, presumably following hydrolysis forming NO2 , is very efficient. Although no absolute

+ rate coefficients were given, the relative rate for reaction of NO2 with phenol as compared to

+ + water k(NO2 + phenol)/k(NO2 + H2O) was determined to be 330 and 1500, for pH of 6 and 10, respectively. The relative rate constants of Heal et al. are similar in magnitude to those found for

+ - + - + 5 reaction of NO2 with Cl as compared to water; k(NO2 + Cl )/k(NO2 + H2O), 450±100 ,

483±1756 and 836±324. When combined, this work implies that nitronium ions may react with

- phenol and Cl with near similar rates. If correct, this suggests that the yield of ClNO2 in dilute chloride matrices may be a function not only of [Cl-], but also the concentration of reactive organics. However, it is important to note that direct reaction rate constants have not been measured in solution.

Recently, Kim et al. measured the flux of N2O5 and ClNO2 at the air-ocean interface at

11 Scripps Institution of Oceanography Pier (La Jolla, CA). It was found that while N2O5 exhibits a net deposition towards the ocean surface during the nighttime hours, ClNO2 was not emitted from the surface as expected based on the modeled reaction kinetics.11 This result is consistent with either: 1) the subsequent reaction or transport of ClNO2 being competitive with its volatilization

+ rate to the gas-phase, or 2) the yield of ClNO2 being suppressed by reaction of NO2 with organics present in the sea surface microlayer (SSML).

Here, we describe a series of laboratory studies designed to measure (ClNO2) following the reaction of N2O5 at the surface of: 1) ambient ocean water samples collected from three locations in the Atlantic and Pacific Ocean, in regions of high and low biological activity, 2) seawater samples collected from a controlled mesocosm experiment conducted in the Scripps

Institution of Oceanography glass-walled wave channel facility, and 3) seawater mimics

86 containing organic proxy molecules representative of those found in SSML. The results are discussed with specific attention to the potential impact that organic molecules, commonly present at atmospheric interfaces, have on ClNO2 production rates.

4.3 Materials and Methods

Yields for ClNO2 production following the reactive uptake of N2O5 were determined in the laboratory by flowing N2O5(g) over both natural and synthetic ocean surfaces contained in

500 mL Pyrex reusable media bottles (Corning Life Sciences) and detecting the resulting

ClNO2(g) using chemical ionization mass spectrometry.

4.3.1 N2O5 Generation and Sample Exposure

Prior to sample exposure, Pyrex media bottles were sequentially rinsed with: acetone,

MilliQ water, and 10-20% HNO3 in MilliQ water solution. Following this, the bottles were dried with UHP N2 to ensure the walls were free of water that could interfere with N2O5 uptake to the liquid surface. The sample solution was transferred directly from the original bottle into a dry 500 mL media bottle using a glass funnel to keep the walls of the headspace free of salt and water.

Solutions were added to the bottles such that the aqueous surface area available for gas-liquid collisions was 47 cm2.

12 N2O5 was generated in situ as described previously by Bertram et al. [2009]. Briefly, ozone is created from photolysis of zero air, utilizing a mercury UV lamp. The ozone flow is then allowed to mix with nitrogen and excess NO2 for approximately 2 minutes. Based the amount of ozone that is titrated following addition of NO2 to the sample stream, we estimate that the N2O5 mixing ratio is 75 ppb. The subsequent 100 sccm N2O5 stream is then passed through a 1/8"

Teflon tube into a Teflon cap fitted with an entrance and exit Swagelok port, which screws onto the top of the sample-containing media bottles. The N2O5 injection line is positioned within 2 cm of the solution surface. All reactions were conducted at room temperature. Based on the volume

87

of the headspace and the sample flow rate, we expect and observe complete removal of N2O5 by the aqueous surface. As such, the experiment does not permit calculation of the N2O5 reaction rate, only (ClNO2).

The 100 sccm product flow was mixed with a wet nitrogen flow after exiting the media bottle, produced by flowing UHP N2 through MilliQ water to ensure high RH. Reaction products were measured using a chemical ionization quadrupole mass spectrometer (CI-QMS), using I- as

- the reagent ion. N2O5 was detected as the I adduct, I·N2O5, as described by Kercher et al., at m/z

13 35 235. Similarly, ClNO2 was detected as I·ClNO2 at m/z 208 for the Cl isotope.

4.3.2 Ambient Sample Collection

Ambient seawater samples were prepared for analysis using the same method as noted above, and stored in 1 L Pyrex media bottles at room temperature until analysis. Ambient seawater samples were collected from three locations. Coastal ocean water samples were collected in La Jolla, CA on two occasions, producing three samples total. During the first sampling period on 14 April 2014, a Pacific Ocean water sample was drawn from the end of

Scripps Institution of Oceanography Pier (La Jolla, CA), and another drawn directly from the shoreline adjacent to Scripps Pier. Similarly, a third sample was collected from shoreline water on

22 July 2014.

Two open ocean water samples were collected from the uncontaminated seawater source onboard the R/V Knorr during the Western Atlantic Cruise II 2014. The first sample was obtained from water with high biological activity on 28 May 2014 (Sample 1, taken at Station 2, 42°

9.88272 N, 61° 40.96026W), while the second was sampled from ocean water with lower biological activity on June 2 2014 (Sample 2, taken at Station 4, 33° 11.38692 N, 63° 9.6855W).

Three seawater samples were taken over the course of a mesocosm phytoplankton bloom produced during the Investigation into Marine PArticle Chemistry and Transfer Science

88

(IMPACTS) campaign 2014 (La Jolla, CA). Recent studies have reported using the same hydraulic facility.14 During IMPACTS 2014, seawater was collected from Scripps Pier, filtered through 50 µm mesh, and added directly into a 33 m × 0.5 m × 1 m glass channel (l × w × h, filled with 9900 L). The wave channel is equipped with a hydraulic paddle capable of generating breaking waves that closely replicate the oceanic bubble entrainment process, and thus the sea spray aerosol distribution of real ocean waves.14,15 The channel was doped with F/2 nutrients

(ProLine F/2 Algae Food, Part# 239800) and sodium metasilicate to encourage the onset of a phytoplankton bloom. The length of the wave channel was fitted with 5700 K fluorescent lights.

Over the course of the mesocosm bloom experiment, three water samples were taken, with one obtained during the pre-bloom period, one during the bloom, and one post-bloom. A suite of water measurements were taken during this campaign in order to characterize the water composition, both biologically and chemically, including chlorophyll-a, dissolved organic carbon, and surface tension measurements of the SSML.

Ambient samples were stored at room temperature for varying amounts of time before analysis. The open ocean sample 1 was stored for 56 days and sample 2 for 51 days. Once obtained, the pre-bloom mesocosm and mid-bloom mesocosm samples were stored in the lab for

11 and 6 days, respectively, before analysis, while the post-bloom mesocosm sample and all coastal samples were obtained and analyzed on the same day.

4.3.3 Laboratory Organic Mimic Preparation

Phenol (J.T. Baker-Avantor Performance Materials, JT-2859-01), humic acid (Spectrum

Chemicals, H1452) and cholesterol (Sigma-Aldrich, C8667) were used as mimics to emulate organic molecules expected to be present in the ocean SSML. Sodium chloride (NaCl) solutions were prepared at standard ocean concentrations (500 mM) using NaCl (Fisher, S271; Sigma

Aldrich, S9888) and Milli-Q water (18.2 MΩ·cm), and mixed thoroughly in a mixing vessel before being transferred to the sampling bottle. Organic mimics were sequentially added to the

89

500 mM NaCl solutions during sample exposure to N2O5. Experiments were run in triplicate, and

the results averaged. Average concentrations used can be found in Table 4.1.

4.3.4 Surface Tension Measurements

Surface tension measurements were taken for ambient seawater samples and laboratory

samples of 500 mM NaCl sequentially doped with phenol. All surface tension measurements

were taken using a Kruss K11 Tensiometer utilizing the Wilhelmy Pt plate method. The glass

sampling cup and Wilhelmy plate were cleaned with acetone and deionized water prior to

measuring each ambient water sample, and with ethanol and deionized water prior to

measurement of the phenol solutions.

Table 4.1 Bulk carbon concentrations following sequential addition of organic molecular mimics to 500 mM NaCl solutions.

Organic Average Bulk [Carbon, (mM)] in solution molecule Following each addition NaCl only Addition 1 Addition 2 Addition 3 Addition 4 Addition 5 Addition 6

Phenol 0 ±0 0.10 ± 0.01 1.00 ± 0.01 5.05 ± 0.01 10.5 ± 1.1 31.6 ± 0.32 48.6 ± 0.49

Humic Acid 0 ±0 0. 23 ± 0.03 0.72 ± 0.08 3.27 ± 0.40 7.26 ± 2.0 NA NA

Cholesterol 0 ±0 0.10 ± 0 1.17 ± 0.06 5.23 ± 0.06 10.3 ± 0.26 NA NA

4.3.5 Box Model

A time-dependent box model was constructed using the standard ordinary differential

equation (ODE) solver package in MatLab to determine theoretical (ClNO2) based on the

chloride and organic concentrations present and the available literature reaction kinetics. Like all

other previous measurements of (ClNO2), this experiment is not capable of directly measuring

90

+ + the aqueous phase rate of NO2 with water, chloride or organic. As such, we set the rate for NO2 with Cl- at the encounter rate (7.4×109 M-1s-1 16) and use literature values of the relative rate

+ coefficient for NO2 reaction with chloride vs. water to calculate (ClNO2) based on chloride, phenol, and water concentrations. In determining the best fit to the data using this model, the sole

+ adjustable parameter is the relative rate constant of NO2 reaction with organics vs. water (i.e.,

+ + k(NO2 + organic)/k(NO2 + H2O)). To the best of our knowledge, neither absolute nor relative

+ rate constants for NO2 reaction with humic acid or cholesterol have been determined at this time.

The constrained model is used to determine the role of organic reactivity in suppressing

(ClNO2).

4.4 Results and Discussion

- 4.4.1 (ClNO2) Dependence on Cl

As shown previously (R4), chloride, acting as a nucleophile, effectively competes with

+ 17,18 H2O for reaction with NO2 . The product yield of ClNO2 following N2O5 hydrolysis at chloride-containing surfaces is directly proportional to the aqueous phase chloride concentration, reaching 100% yield when chloride concentrations are greater than 4.6 M. We first replicated the

- ClNO2 yield curve as a function of Cl to confirm that our experimental approach was consistent with that of previous studies. Figure 4.1 shows the N2O5 (panel A) and ClNO2 (panel B) signals for N2O5 flowing to the CI-QMS inlet (section I), over a saturated NaCl solution (6.1 M NaCl)

(section II) and back to the instrument inlet (section III). As shown, all N2O5 is lost when flowing over the saturated NaCl solution, and subsequently ClNO2 is recovered. Since this solution is saturated, the conversion of N2O5 to ClNO2 is taken to be 100% based the studies of Bertram and

Thornton, Behnke et al., and Roberts et al..4–6 Calibration is required, as the CI-QMS sensitivity to N2O5 and ClNO2 is not equivalent, as indicated by the signal levels shown in Figure 4.1.

91

Figure 4.1 Signal intensities for N2O5 (top panel) and ClNO2 (bottom panel) during N2O5 delivery to CI- QMS inlet (I), after flowing over a saturated NaCl solution (6.1 M NaCl) (II), and back to the CI-QMS inlet (III).

The (ClNO2) curve as determined using our experimental approach is shown in Figure

+ - + 4.2, alongside literature results. We calculate k(NO2 + Cl )/k(NO2 + H2O) to be in general agreement with the results of previous studies measuring the relative rate constants of the two

+ - + 5 6 4 reactions as k(NO2 + Cl )/k(NO2 + H2O) between 450±100 , 483±175 and 836±32 , providing

92

confidence that our experimental approach is an accurate method for measuring (ClNO2) from aqueous interfaces.

During each experiment conducted using ambient seawater or organic mimics, either a saturated solution of NaCl or a 500 mM NaCl solution was used as an internal calibration for

(ClNO2). When an internal calibration was not available, the relative sensitivity of the quadrupole mass spectrometer towards ClNO2 and N2O5 was used to calculate (ClNO2). The calculated ClNO2 yields were invariant to the reference method, within the reported uncertainty.

Figure 4.2 ClNO2 yield as a function of chloride concentration as determined in this work (red squares) and fit values from previous determinations, Roberts et al., 2009 (circles), Bertram and Thornton, 2009 (inverted triangles), and Behnke et al., 1997 (diamonds).

4.4.2 ClNO2 Yields from Ambient Seawater Samples

The ambient seawater samples chosen provide a spatially diverse set of samples permitting assessment of the influence of seawater organic composition and concentration on

93

ClNO2 product yields. Table 4.2 shows the retrieved (ClNO2) for each ambient water sample, along with the co-located chlorophyll-a concentrations, an indicator of biological activity, and any additional water side measurements. Based on the (ClNO2) curve from this work, shown in

- Figure 4.2, we expect a ClNO2 yield of 0.82 ± 0.05 based on a seawater Cl concentration of 500 mM.

Coastal seawater samples were collected from the Scripps Institution of Oceanography

(SIO) Pier during times of both high and low chlorophyll-a concentration with the intent of probing how (ClNO2) is impacted by a change in biological activity in the surface waters, inferred from chlorophyll-a concentration. Surprisingly, (ClNO2) obtained from the seawater samples were not statistically different from one another, but were both significantly suppressed as compared with that of a NaCl mixture of the same chloride concentration (0.82 ± 0.05 for 500 mM Cl-). The two seawater samples taken during a period of elevated chlorophyll-a (4.66 ± 0.37

µg/L) resulted in ClNO2 yields of 0.23 ± 0.18 and 0.21 ± 0.19 for shore and pier water, respectively. The single seawater sample taken at a chlorophyll-a concentration of 2.00 ± 0.21

µg/L resulted in a ClNO2 yield of 0.30 ± 0.08.

94

Table 4.2 Ambient water samples, corresponding ClNO2 yields, and co-located measurements of chlorophyll-a, colored dissolved organic matter (CDOM), dissolved organic carbon (DOC), and surface tension where available.

*5 hour average and standard deviation calculated from SCCOOS data set.

Sample Location Measured Chl-a (µg/L) CDOM (ppb) DOC(M, Surface Tension Surface Tension

ClNO2 Yield %CV) in Bulk (mN/m) in SSML (mN/m)

Mesocosm sample 1 0.26 ± 0.02 0.92 ± 0.09 3.76 ± 0.48 233.46, 1.51 73.1 ± 0.4 72.2 ± 0.3 Mesocosm sample 2 0.24 ± 0.05 2.43 ± 0.03 2.95 ±0.41 245.96, 1.70 72.6 ± 0.7 72.8 ± 0.4 Mesocosm sample 3 0.16 ± 0.05 0.74 ± 0.16 2.93 ± 0.26 237.52, 0.77 73.5 ± 0.1 73.1 ± 1.2 Coastal water, SIO 0.23 ± 0.18 4.66 ± 0.37* NA NA NA NA pier, shore sample 1 Coastal water, SIO 0.21 ± 0.19 4.66 ± 0.37* NA NA NA NA pier, pier sample Coastal water, SIO 0.30 ± 0.08 2.00 ± 0.21* NA NA NA NA pier, shore sample 2 Open ocean, sample 1 0.27 ± 0.03 1.96 ± 1.00 NA NA NA NA

Open ocean, sample 2 0.20 ± 0.06 0.08 ± 0.02 NA NA NA NA

For comparison, seawater from the Atlantic Ocean was obtained in two locations, one from an area of high biological activity (chlorophyll-a value of 1.96 ± 1.00 µg/L), and one of low biological activity (chlorophyll-a concentration of 0.08 ± 0.02 µg/L). Again, surprisingly, the retrieved (ClNO2) was comparable for both samples (0.27 ± 0.03 and 0.20 ± 0.06) and significantly lower than that determined from the NaCl synthetic mixture at the same chloride concentration.

To investigate the effects of organic content in more detail, seawater samples were collected at three different points during a controlled mesocosm bloom experiment as part of the

2014 IMPACTS campaign. Over the course of a microbial bloom in the ocean, there are generally stages during which different microorganisms dominate. This cycle begins with a predominance of phytoplankton, followed by either their natural decay or decay via bacterial consumption, at which point bacteria dominate as they prey on either living or decayed phytoplankton, and finally viruses dominate after feeding on the bacteria and other living cells in

95 the ocean.19–22 During each step of the microbial loop, dissolved organic matter is released into the water. Due to the predominance of different microbes during each step of the cycle, it is expected that the water composition, specifically with regards to organic chemical composition and concentration, will vary significantly.

Based on chlorophyll-a measured during IMPACTS 2014, three water samples were collected that fall in pre-bloom (chlorophyll-a = 0.92 ± 0.09 µg/L), peak bloom (chlorophyll-a =

2.43 ± 0.03 µg/L), and post bloom (chlorophyll-a = 0.74 ± 0.16 µg/L) periods. At the points in which samples were taken during this study, values for DOC were 0.23 ± 3.5×10-3 , 0.25 ±

4.2×10-3 , and 0.24 ± 1.8 ×10-3 mM C for the three sampling periods respectively, while the values for SSML surface tension were 72.2 ± 0.3, 72.8 ± 0.4, and 73.1 ± 1.2 mN/m, respectively.

Consistent with the coastal ocean water samples, (ClNO2) is suppressed, and remarkably insensitive to ocean conditions with yield values of 0.26 ± 0.02 (pre-bloom), 0.24 ± 0.05 (mid- bloom), and 0.16 ± 0.05 (post-bloom). Figure 4.3 shows (ClNO2) measured as a function of chlorophyll-a concentration for all of the ambient seawater measurements made in this study.

Surprisingly, (ClNO2) as determined for seawater is both significantly suppressed as compared to (ClNO2) for an equivalent NaCl solution, but is also remarkably consistent, ranging between

0.16 - 0.30. This suggests that the mechanism by which ClNO2 production is suppressed is relatively constant, even over the course of a large-scale phytoplankton bloom.

4.4.3 (ClNO2) from Laboratory Mimics

A series of controlled laboratory experiments were run in order to understand the role of organics in limiting the production of ClNO2, following N2O5 hydrolysis at the ocean surface.

The observation of a suppressed, but invariant (ClNO2) suggests that the refractory component of dissolved organic matter may control (ClNO2) at the atmosphere-ocean interface. To test this hypothesis, we selected phenol, humic acid, and cholesterol as potential mimics of the refractory

96 component of the sea-surface microlayer. Previous work has shown that aromatic carbon, especially in the form of fulvic and humic-like substances, is present in the sea surface microlayer.23,24,25

Figure 4.3 Dependence of the measured ClNO2 yield on chlorophyll-a, a proxy for biological activity, for ambient sea-water samples collected from coastal and open ocean water and from a large scale mesocosm bloom experiment.

The dependence of (ClNO2) on phenol was explored initially, as there is existing

+ 10 literature on the relative rate constant of NO2 with phenol as compared to water that can be

+ combined with the relative rate constants of NO2 with chloride and with water to assess the potential effect of aromatics on (ClNO2). While it has not been experimentally verified, laboratory measurements are most consistent with a model that assumes that N2O5 hydrolysis occurs well within the first 50 nm of the interface and is thus sensitive to near-interfacial rather than bulk concentrations.26,27 We proceed with a model estimate of the phenol and Cl- interfacial

97 concentrations by assuming that Cl- depletion and phenol segregation are limited to the outermost single layer of the solution. To determine the surface phenol:chloride ratio, we first calculate the

2-dimensional (2D) concentration of phenol and chloride assuming no surface enhancement or depletion of either species, which can be approximated by Equation 1,28

E1

where cspecies is the bulk concentration of solute (variable in the case of phenol (0 - 167 mM and

500 mM for NaCl), cwater is the concentration of water in solution (55.5 M), Aspecies is the cross

-15 2 29 - -15 2 30 sectional area of the molecule (phenol = 3.0×10 cm , Cl = 1.0×10 cm ), and Awater is the cross sectional area of a water molecule (1.0×10-15 cm2). The 2D concentrations calculated for phenol varied based on the concentration of phenol in solution, but were between 3.0×108 -

3.0×1012 molecules/cm2, while the 2D concentration obtained for chloride at 500 mM was

8.9×1012 molecules/cm2.

Chloride is known to be present in lower concentrations at the air-water interface as compared to the bulk.31,32 Interfacial chloride depletion is accounted for through calculation of surface depletion based on measurements of surface tension. Similarly, we expect an enrichment of phenol at the surface, thus requiring a calculation of the phenol surface excess. Surface tension measurements were made on 500 mM NaCl solutions for phenol bulk concentrations between 0 -

167 mM, with results for 0 - 8.1 mM shown in Figure 4.4A. The increase in surface tension with increasing NaCl in solution is well known, thus surface tension values from the literature were used to calculate the Cl- 2D concentration.33

98

Figure 4.4 Surface tension of laboratory samples as a function of [phenol]bulk (A, red squares) along with a linear fit to the data (dashed line). Calculated phenol surface excess (B, blue circles), as derived from the linear fit shown in Panel A, and the calculated concentration of chloride at the solution interface (dashed line). The molar ratio of phenol to chloride at the surface (left axis) and in the bulk (right axis), derived from the slope of the linear fit in Panel A and [phenol]bulk values is shown in panel C. The dashed line is a linear fit.

For the relatively dilute concentrations used here, the change in activity coefficient is small, and thus the surface excess (or depletion) in phenol (or chloride) was calculated using the following form of the Gibbs equation,34,35 given by Equation 2:

99

E2

where Γ is the surface excess, R is the molar gas constant, T is temperature,  is surface tension, and c is the bulk concentration of the solution. For phenol, /lnc was obtained from the fit to the data in Figure 4.4A. For 500 mM Cl- in solution, the surface excess value calculated was -

8.3×1012 cm-2, in good agreement with previous literature values.36 The phenol surface excess values obtained for bulk concentrations ranging from 0.02 - 167 mM were 5.5×1010 to 5.6×1014 molecules/cm2, in agreement with results by Rao et al..37 The resulting surface excess (or depletion in the case of Cl-) for each concentration is combined with the respective 2D surface concentration calculated via Equation 1, to determine the total species concentration at the surface, assuming that deviations from the bulk are confined to the outermost monolayer

(E3),28,35,38

E3

The resulting monolayer concentrations for phenol are plotted in Figure 4.4B versus the bulk phenol concentration. The dashed line in Figure 4.4B shows the total surface concentration of Cl- (6.8×1011 molecules/cm2), assuming that the interfacial Cl- concentration is not altered by phenol segregation. The phenol surface concentration reaches a maximum of ~2×1013 molecules/cm2, where it covers roughly 6% of the surface. Figure 4.4B highlights that the equivalence point, where phenol and chloride are present in equal concentrations at the interface, occurs at very low bulk phenol:Cl- ratios, which in this model is roughly 1:2000. This small ratio implies that the surface composition can be significantly different than the bulk composition for

100 very subtle decreases in surface tension (73.4 to 72.1 mN m-1). This is further illustrated in Figure

4.4C where the ratio of phenol to Cl- at the surface is plotted vs. the bulk phenol concentration.

- Here, for a [phenol]bulk of 8.1 mM, the bulk ratio of phenol:Cl is 1:63, while the surface ratio of phenol:Cl- is greater than 39:1.

As shown in Figure 4.5, (ClNO2) decreases with increasing phenol concentrations, as expected, down to 0.33 for a phenol:Cl- surface ratio of 39:1 (Figure 4.5). Employing the box

+ + - model discussed previously, we fit the relative rate of k(NO2 + organic)/k(NO2 + Cl ) to best match the observed (ClNO2) using the calculated surface concentrations. The model results are

+ shown as lines in Figure 4.5. Here, we bracket the measured data using the relative rate of NO2

- + + - with phenol compared to that with Cl (i.e., k(NO2 + organic) /k(NO2 + Cl )). It is found that the

+ + - data is most consistent with k(NO2 + organic) of 0.04-0.15 times that of k(NO2 + Cl )).

The effect of more ocean relevant, insoluble non-aromatic and aromatic organics was probed using cholesterol (Figure 4.6, blue squares), and humic acid (Figure 4.6, black circles), respectively. The concentrations shown in Figure 4.6 assume homogeneous mixing throughout the solution. However, due to the low solubility of humic acid and cholesterol, we expect that the concentration in the reaction region at the surface is significantly enhanced as compared to the bulk, as was shown for phenol. Unfortunately, we do not have robust surface tension measurements for these samples. The addition of cholesterol results in no notable reduction in

ClNO2 yield. In contrast, humic acid addition results in a significant reduction in (ClNO2) to values as low as 0.24 ± 0.09 (Figure 4.6) assuming a homogenous bulk carbon concentration of

10 mM in the water column. However, in the case of humic acid, the assumption of surface enrichment must be considered since the solubility of humic acid in water is minimal, thus producing an organic concentration gradient throughout the water column. Additionally, as noted

26 above, the reacto-diffusive length for N2O5 has been shown to be less than 50 nm at 50 % RH, thus we can assume that the surface concentration enhancement is influencing (ClNO2)

101 suppression, making it appear that the nitronium ion is reacting with humic acid at a rate faster than the collision limit.

4.5 Linking Laboratory and Field Studies.

+ It has been shown previously that the nitration of phenol by NO2 is of order 10000 times faster than benzene in perchloric acid solutions of varying acidities8 and anisole up to 15000 times faster than benzene in solutions of acetic anhydride7, indicating that the rate of nitration is a function of the degree of aromaticity and the nature of any electron donating functional groups, as might be expected. This has implications for understanding the rate of reaction of complex species, such as humic or fulvic acids on the surface of the ocean.

Figure 4.5 ClNO2 yield determinations (red squares) as a function of the molar ratio of phenol to chloride at the surface (bottom axis) and in the bulk (top axis) following sequential additions of phenol to 500 mM NaCl solution. The solid lines represent model results, which bracket the observed data points, obtained by + + - varying k(NO2 + Phenol) with respect to k(NO2 + Cl ), which is assumed to proceed at the encounter rate.

102

Our results suggest that in order to explain the observed suppression in (ClNO2) for

+ seawater samples, the concentration of highly reactive organics (k ≥ k(NO2 + phenol)) needs to exceed the concentration of Cl- at the interface by more than a factor of 40 (Figure 4.5). As shown above, this condition can be met with relatively small changes in surface tension which may exist for ocean samples. While it has been shown that the carbon concentration in the ocean ranges between 0.034-0.090 mM,39,40considerably less is known regarding the concentration and speciation of organics in the SSML, operationally defined as the top 60-1000 µm of the ocean surface.41–43 This rich organic matrix provides the interface between the ocean and atmosphere and, due to its composition, can exhibit differing behavior compared to bulk seawater. The enhancement in dissolved organic matter has been shown to be the result of organic scavenging as bubbles rise through the water column in the ocean.43,44 Previous studies have determined this organic-rich region exhibits a 15-25 times enrichment of viruses and bacteria,45,46 a 50 times enhancement in amino acids,47 along with up to a 10 times enrichment of dissolved organic compounds when compared to corresponding bulk sub waters.48,49 For example, based on a bulk dissolved organic carbon (DOC) concentration of 0.070 mM,39,40 DOC concentrations in the

SSML could be as high as 0.7 mM. It should be noted that typical methods of SSML sampling tend to interrogate the top 50-400 m of the ocean surface.45,50 If the stratification of organics in the ocean continues throughout the SSML, it is possible that the top 50 nm of the surface, most relevant to N2O5 reactive uptake given the reacto-diffusive length of N2O5 is <50 nm at 50%

RH,26 could have organic concentrations enhanced significantly above the values given above.

Future work focused on high precision measurements of surface tension for ambient samples may help bridge these gaps.

103

Figure 4.6 ClNO2 yield as a function of carbon concentration, assuming homogenous mixing, following sequential additions of cholesterol (blue squares) and humic acid (black circles) into a 500 mM NaCl solution.

4.6 Summary and Conclusions

Three primary observations from this study are: 1) seawater samples obtained from both coastal and open ocean waters of differing biological conditions all show a marked reduction in

(ClNO2), ranging between 16 and 30%, regardless of chlorophyll-a concentration, 2) aromatic organics, known to be present in the SSML, are capable of reducing (ClNO2) following the reactive uptake of N2O5 to the aqueous surface, and 3) relatively small changes in surface tension can reflect dramatic enhancements in the surface organic:Cl- ratio, which in turn serves to suppress ClNO2 to levels unexpected based on bulk concentrations. Collectively, these results provide further confidence that N2O5 hydrolysis occurs in the top tens of nanometers from the interface, a region where organic:Cl- ratios are enhanced relative to the bulk. These observations suggest that measurement of (ClNO2) at an aqueous interface may serve as an effective probe of the SSML chemical composition and that aromatic organics present in the sea surface microlayer

104

have the potential to contribute to the suppressed production of ClNO2 following nocturnal N2O5 uptake to the ocean surface.

- + While our results are consistent with a competition between Cl and phenol for NO2 , it is also possible that the suppression in (ClNO2) may result from the direct reaction of ClNO2 with organic molecules such as phenol. Our model assumes prompt ClNO2 volatilization following production in the aqueous phase. If ClNO2 hydrolysis is instead fast compared with volatilization,

+ the subsequent reaction of this NO2 to form stable, non-volatile products (e.g., organic nitrates), would be interpreted as a suppression in (ClNO2). In order to further understand the

+ contribution of the NO2 -unsaturated organic reaction rate and organic concentration to (ClNO2) suppression in the SSML, it is necessary that future measurements focus on direct determinations

+ - of the reaction rate of NO2 with H2O, Cl and various organic molecules. Further, simultaneous measurements of aqueous phase reaction products, e.g. nitrated and/or chlorinated aromatics will provide unique insight on the reaction mechanism.

4.7 Acknowledgements

Chapter 4, in full, is currently under review in The Journal of Physical Chemistry A.

O.S. Ryder, N.R. Campbell1, M. Shaloski, H. Al-Mashat, G. M. Nathanson, and T.H. Bertram.

(2014), On the Role of Organics in Regulating ClNO2 Production at the Air-Sea Interface. The dissertation author was the primary investigator and author of this material.

This research was supported by the National Science Foundation via the Center for

Aerosol Impact on Climate and the Environment, a Center for Chemical Innovation (NSF

CHE1305427). The authors gratefully acknowledge Dr. Grant Deane and Dale Stokes for use of the tensiometer used in this work. Dr. Matthew Zoerb (UCSD) for the collection of Western

Atlantic Cruise II 2014 water samples, Bob Vaillancourt, Jeremiah Stone, and Evan Ntonados,

(all from Millersville University of Pennsylvania) for providing WACS2 chlorophyll-a

105 concentrations, and the entire CAICE IMPACTS team and the staff of the SIO Hydraulics

Laboratory. O.S.R gratefully acknowledges a Graduate Research Fellowship from the National

Science Foundation (2011-2014).

4.8 References

(1) Abbatt, J. P. D.; Lee, A. K. Y.; Thornton, J. A. Quantifying Trace Gas Uptake to Tropospheric Aerosol: Recent Advances and Remaining Challenges. Chem. Soc. Rev. 2012, 41 (19), 6555–6581.

(2) Brown, S. S.; Stark, H.; Ravishankara, A. R. Applicability of the Steady State Approximation to the Interpretation of Atmospheric Observations of NO3 and N2O5. J. Geophys. Res. Atmospheres 2003, 108 (D17), 4539.

(3) Wayne, R. P.; Barnes, I.; Biggs, P.; Burrows, J. P.; Canosa-Mas, C. E.; Hjorth, J.; Le Bras, G.; Moortgat, G. K.; Perner, D.; Poulet, G.; Restelli, G.; Sidebottom, H. The Nitrate Radical: Physics, Chemistry, and the Atmosphere. Atmospheric Environ. Part Gen. Top. 1991, 25 (1), 1–203.

(4) Behnke, W.; George, C.; Scheer, V.; Zetzsch, C. Production and Decay of ClNO2 from the Reaction of Gaseous N2O5 with NaCl Solution: Bulk and Aerosol Experiments. J. Geophys. Res. Atmospheres 1997, 102 (D3), 3795–3804.

(5) Bertram, T. H.; Thornton, J. A. Toward a General Parameterization of N2O5 Reactivity on Aqueous Particles: The Competing Effects of Particle Liquid Water, Nitrate and Chloride. Atmos Chem Phys 2009, 9 (21), 8351–8363.

(6) Roberts, J. M.; Osthoff, H. D.; Brown, S. S.; Ravishankara, A. R.; Coffman, D.; Quinn, P.; Bates, T. Laboratory Studies of Products of N2O5 Uptake on Cl− Containing Substrates. Geophys. Res. Lett. 2009, 36 (20), L20808.

(7) Schofield, K. Aromatic Nitration; Cambridge University Press: Great Britain, 1980.

(8) Hoggett, J. G.; Moodie, R. B.; Penton, J. R.; Schofield, K. Nitration and Aromatic Reactivity; Cambridge University Press: Great Britain, 1971.

(9) Robert Grossman. The Art of Writing a Reasonable Organic Reaction Mechanism, 2nd ed.; Springer, 2003.

(10) Heal, M. R.; Harrison, M. A. J.; Neil Cape, J. Aqueous-Phase Nitration of Phenol by N2O5 and ClNO2. Atmos. Environ. 2007, 41 (17), 3515–3520.

(11) Kim, M. J.; Farmer, D. K.; Bertram, T. H. A Controlling Role for the Air−sea Interface in the Chemical Processing of Reactive Nitrogen in the Coastal Marine Boundary Layer. Proc. Natl. Acad. Sci. 2014, 111 (11), 3943–3948.

106

(12) Bertram, T. H.; Thornton, J. A.; Riedel, T. P. An Experimental Technique for the Direct Measurement of N2O5 Reactivity on Ambient Particles. Atmos Meas Tech 2009, 2 (1), 231–242.

(13) J. P. Kercher, T. P. R. Chlorine Activation by N2O5: Simultaneous, in Situ Detection of ClNO2 and N2O5 by Chemical Ionization Mass Spectrometry. Atmospheric Meas. Tech. 2009, 2 (1).

(14) Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; DeMott, P. J.; Aluwihare, L. I.; Palenik, B. P.; Azam, F.; Seinfeld, J. H.; Moffet, R. C.; Molina, M. J.; Cappa, C. D.; Geiger, F. M.; Roberts, G. C.; Russell, L. M.; Ault, A. P.; Baltrusaitis, J.; Collins, D. B.; Corrigan, C. E.; Cuadra-Rodriguez, L. A.; Ebben, C. J.; Forestieri, S. D.; Guasco, T. L.; Hersey, S. P.; Kim, M. J.; Lambert, W. F.; Modini, R. L.; Mui, W.; Pedler, B. E.; Ruppel, M. J.; Ryder, O. S.; Schoepp, N. G.; Sullivan, R. C.; Zhao, D. Bringing the Ocean into the Laboratory to Probe the Chemical Complexity of Sea Spray Aerosol. Proc. Natl. Acad. Sci. 2013, 110 (19), 7550–7555.

(15) Deane, G. B.; Stokes, M. D. Scale Dependence of Bubble Creation Mechanisms in Breaking Waves. Nature 2002, 418 (6900), 839–844.

(16) Mortimer, R. Physical Chemistry, 2nd ed.; Harcourt Academic Press, 2000.

(17) Finlayson-Pitts, B. J.; Ezell, M. J.; Pitts, J. N. Formation of Chemically Active Chlorine Compounds by Reactions of Atmospheric NaCl Particles with Gaseous N2O5 and ClONO2. Nature 1989, 337 (6204), 241–244.

(18) Francis Schweitzer, P. M. Multiphase Chemistry of N2O5, ClNO2, and BrNO2. J. Phys. Chem. A 1998, 102 (22).

(19) Russell, L. M.; Hawkins, L. N.; Frossard, A. A.; Quinn, P. K.; Bates, T. S. Carbohydrate- like Composition of Submicron Atmospheric Particles and Their Production from Ocean Bubble Bursting. Proc. Natl. Acad. Sci. 2010, 107 (15), 6652–6657.

(20) Quinn, P. K.; Bates, T. S. The Case against Climate Regulation via Oceanic Phytoplankton Sulphur Emissions. Nature 2011, 480 (7375), 51–56.

(21) Pomeroy, L.; leB. Williams, P.; Azam, F.; Hobbie, J. The Microbial Loop. Oceanography 2007, 20 (2), 28–33.

(22) Azam, F.; Malfatti, F. Microbial Structuring of Marine Ecosystems. Nat. Rev. Microbiol. 2007, 5 (10), 782–791.

(23) Tschapek, M.; Wasowski, C.; Sánchez, R. M. T. Humic Acid as a Colloidal Surfactant. Plant Soil 1981, 63 (2), 261–271.

(24) Harvey, G. R.; Boran, D. A.; Chesal, L. A.; Tokar, J. M. The Structure of Marine Fulvic and Humic Acids. Mar. Chem. 1983, 12 (2–3), 119–132.

107

(25) Carlson, D. J. Surface Microlayer Phenolic Enrichments Indicate Sea Surface Slicks. Nature 1982, 296 (5856), 426–429.

(26) Thornton, J. A.; Braban, C. F.; Abbatt, J. P. D. N2O5 Hydrolysis on Sub-Micron Organic Aerosols: The Effect of Relative Humidity, Particle Phase, and Particle Size. Phys. Chem. Chem. Phys. 2003, 5 (20), 4593–4603.

(27) Stewart, D. J.; Griffiths, P. T.; Cox, R. A. Reactive Uptake Coefficients for Heterogeneous Reaction of N2O5 with Submicron Aerosols of NaCl and Natural Sea Salt. Atmos. Chem. Phys. 2004, 4 (5), 1381–1388.

(28) DeZwaan, J. L.; Brastad, S. M.; Nathanson, G. M. The Roles of Salt Concentration and Cation Charge in Collisions of Ar and DCl with Salty Glycerol Solutions of NaI and CaI2. J. Phys. Chem. C 2008, 112 (8), 3008–3017.

(29) Hórvölgyi, Z.; Kiss, E. Colloids for Nano- and Biotechnology; Springer Science & Business Media, 2008.

(30) Marcus, Y. Ionic Radii in Aqueous Solutions. Chem. Rev. 1988, 88 (8), 1475–1498. (31) Cummings, O. T.; Wick, C. D. Interfacial Behavior of Simple Inorganic Salts at the Air- Water Interface Investigated with a Polarizable Model with Electrostatic Damping. J. Chem. Phys. 2013, 139 (6), 064708.

(32) Jungwirth, P.; Tobias, D. J. Molecular Structure of Salt Solutions: A New View of the Interface with Implications for Heterogeneous Atmospheric Chemistry. J. Phys. Chem. B 2001, 105 (43), 10468–10472.

(33) Ozdemir, O.; Karakashev, S. I.; Nguyen, A. V.; Miller, J. D. Adsorption and Surface Tension Analysis of Concentrated Alkali Halide Brine Solutions. Miner. Eng. 2009, 22 (3), 263–271.

(34) Jungwirth, P.; Tobias, D. J. Specific Ion Effects at the Air/Water Interface. Chem. Rev. 2006, 106 (4), 1259–1281.

(35) Chattoraj, D. K.; Birdi, K. S. Adsorption at Liquid Interfaces and the Gibbs Equation. In Adsorption and the Gibbs Surface Excess; Springer US, 1984; pp 39–82.

(36) Shah, A.-H. A.; Ali, K.; Bilal, S. Surface Tension, Surface Excess Concentration, Enthalpy and Entropy of Surface Formation of Aqueous Salt Solutions. Colloids Surf. Physicochem. Eng. Asp. 2013, 417, 183–190.

(37) Rao, Y.; Subir, M.; McArthur, E. A.; Turro, N. J.; Eisenthal, K. B. Organic Ions at the Air/water Interface. Chem. Phys. Lett. 2009, 477 (4–6), 241–244.

(38) Butler, J. a. V.; Wightman, A. 293. Adsorption at the Surface of Solutions. Part I. The Surface Composition of Water–alcohol Solutions. J. Chem. Soc. Resumed 1932, No. 0, 2089–2097.

108

(39) Hansell, D.; Carlson, C.; Repeta, D.; Schlitzer, R. Dissolved Organic Matter in the Ocean: A Controversy Stimulates New Insights. Oceanography 2009, 22 (4), 202–211.

(40) Ogawa, H.; Tanoue, E. Dissolved Organic Matter in Oceanic Waters. J. Oceanogr. 2003, 59 (2), 129–147.

(41) Wurl, O.; Obbard, J. P. A Review of Pollutants in the Sea-Surface Microlayer (SML): A Unique Habitat for Marine Organisms. Mar. Pollut. Bull. 2004, 48 (11–12), 1016–1030.

(42) Cunliffe, M.; Engel, A.; Frka, S.; Gašparović, B.; Guitart, C.; Murrell, J. C.; Salter, M.; Stolle, C.; Upstill-Goddard, R.; Wurl, O. Sea Surface Microlayers: A Unified Physicochemical and Biological Perspective of the Air–ocean Interface. Prog. Oceanogr. 2013, 109, 104–116.

(43) Hardy, J. T. The Sea Surface Microlayer: Biology, Chemistry and Anthropogenic Enrichment. Prog. Oceanogr. 1982, 11 (4), 307–328.

(44) Batoosingh, E.; Riley, G. A.; Keshwar, B. An Analysis of Experimental Methods for Producing Particulate Organic Matter in Sea Water by Bubbling. Deep Sea Res. Oceanogr. Abstr. 1969, 16 (2), 213–219.

(45) Aller, J. Y.; Kuznetsova, M. R.; Jahns, C. J.; Kemp, P. F. The Sea Surface Microlayer as a Source of Viral and Bacterial Enrichment in Marine Aerosols. J. Aerosol Sci. 2005, 36 (5– 6), 801–812.

(46) Blanchard, D. C.; Syzdek, L. Mechanism for the Water-to-Air Transfer and Concentration of Bacteria. Science 1970, 170 (3958), 626–628.

(47) Kuznetsova, M.; Lee, C. Dissolved Free and Combined Amino Acids in Nearshore Seawater, Sea Surface Microlayers and Foams: Influence of Extracellular Hydrolysis. Aquat. Sci. 2002, 64 (3), 252–268.

(48) Thurman, E. M. Organic of Natural Waters; Kluwer Academic Publishers: USA, 1985.

(49) Garabetian, F.; Romano, J.-C.; Paul, R.; Sigoillot, J.-C. Organic Matter Composition and Pollutant Enrichment of Sea Surface Microlayer inside and Outside Slicks. Mar. Environ. Res. 1993, 35 (4), 323–339.

(50) Van Pinxteren, M.; Müller, C.; Iinuma, Y.; Stolle, C.; Herrmann, H. Chemical Characterization of Dissolved Organic Compounds from Coastal Sea Surface Microlayers (Baltic Sea, Germany). Environ. Sci. Technol. 2012, 46 (19), 10455–10462.

Chapter 5 The potential role of divalent cations at the air-sea interface in regulating the production rate of ClNO2

5.1 Abstract

The reaction of N2O5 at aqueous chloride-containing atmospheric interfaces, including the ocean surface and aerosol particles, impacts the lifetime of reactive nitrogen and the production rate of chlorine radicals. Recent field observations have shown that N2O5 deposits to

1 the ocean surface with no surface resistance. However, corresponding emission of ClNO2 from the interface has not been observed.1 Laboratory studies using ambient seawater collected from various locations (including the Atlantic and Pacific Ocean) and under varied biological conditions (chlorophyll-a = 0.08-4.66 µg/L) also indicate that the production rate of ClNO2 is suppressed, where the ClNO2 yield was measured to be between 0.16 - 0.30 for all seawater samples,2 significantly smaller than that expected for a 500mM Cl- interfaces (0.82 ± 0.05).

- Laboratory measurements of the ClNO2 yield, conducted at Cl concentrations typical of ocean conditions ([Cl-] = 500mM), were capable of replicating the yield observe for the ambient samples, but only upon addition of reactive, surface active organics such as phenol. Here, we

2+ 2+ explore the potential role for divalent cations (e.g., Mg and Ca ) to suppress the yield of ClNO2 at the air-sea interface, through modulating the availability of Cl- at the air-sea interface. This work has implications for both reactive halogen and nitrogen budgets in polluted marine environments.

5.2 Introduction

The work in this chapter builds upon the work presented in Chapter 4, which focused on determinations of ClNO2 yields obtained after flowing N2O5 over the surface of ambient seawater collected from a coastal location (Pacific Ocean) and from an open ocean location (Atlantic

109

110

Ocean), and during a stimulated phytoplankton bloom in the laboratory using real seawater.2

Multiple water samples were collected from each of these locations, and exhibited different degrees of biological activity, as indicated by chlorophyll-a concentrations. The yield of ClNO2,

(ClNO2) obtained from all samples was consistently suppressed from the expected value of 0.82

± 0.05, down to 0.16 - 0.30 irrespective of biological conditions in the ocean. The work presented in Chapter 4 indicates that reactive aromatic species, such as phenol and humic acid, exhibit a concentration gradient within the top surface of the ocean, and are capable of suppressing

(ClNO2) down to levels observed in real ocean water samples. However, the results also indicate that to achieve the suppression observed in the ambient samples the molar ratio of reactive organic to chloride must exceed 2 in the reaction region (top 1nm of the interface). While it is likely that the organic reaction provides a component of the observed suppression in ClNO2, it is possible that trace divalent cations that are present in seawater may also contribute to the mechanism of reducing chloride concentrations at the interface.

The bulk concentration of organics in seawater is typically 0.034-0.090 mM,3,4 with an average value often cited as 0.070 mM, and with little known about the speciation of such material. The inorganic components of sea water, however, are better understood in terms of identity, concentration, and the variability in concentration is minimal across the oceans.5,6 In

3.5% salinity seawater, the principal ions are chloride (1.89×104 ppm), sodium (1.06×104 ppm), sulfate (2.65×103 ppm), magnesium (1.27×103 ppm), calcium (400 ppm), and potassium (380 ppm), bicarbonate (140 ppm) and bromide, borate, strontium, and fluoride all <100ppm.7 While it is known that the salinity in the oceans can vary within a small range (3.3 to 3.7 %) based on evaporation and precipitation rates in a given region, the relative ratio of these principal ions to one another remains constant.6

Previous theoretical work has shown that inorganic ions such as Mg2+, Cl-, and F- are present at reduced concentrations at the air-liquid interface.8,9Additionally, Mg2+, has been shown

111

- + to affect the microsolvation of Cl at the interface, potentially impacting nitronium ion (NO2 )

10,11 reaction and thus (ClNO2).

Given the consistency in the relative abundances of inorganic ions across the oceans, and its invariance with biological activity, it is possible an inorganic species may either directly or indirectly influence the yield of ClNO2 from the ocean surface.

5.3 Materials and Methods

5.3.1 Solution Preparation

Glassware was cleaned prior to sample preparation as previously described in Chapter 4.2

Briefly, Pyrex media bottles were sequentially rinsed with acetone, MilliQ water, and 10-20%

HNO3 in MilliQ water. The bottles were then dried using a dry flow of nitrogen to ensure the walls were free of any water that could affect the uptake of N2O5. Solutions of 500 mM NaCl were prepared and transferred to a clean, dry 500 mL media bottle using a long spouted glass funnel to ensure the solution did not contact the walls of the headspace. Solutions were added so

2 that the liquid surface area available for N2O5 reaction was 47 cm .

5.3.2 Real and Proxy Ocean Water Solutions

A series of salt solutions were studied in these experiment. Artificial reef salt, a mixture commercially available for aquarium use, made from ACS-grade chemicals (Neomarine Precision

Salt Blend for Reef Aquaria, Brightwell Aquatics), was used as a proxy for sea salt. For all experiments utilizing this salt, the salt was prepared to a salinity of 3.5%.

Sodium chloride solutions ([Cl-] = 500mM, Sigma Aldrich, ≥ 99.0%), were doped with, either MgCl2 .6H2O (Either Sigma Aldrich, 99.0-102.0 % ACS reagent grade or Macron Fine

Chemicals, 99.0-102.0 % ACS reagent grade), MgCl2 anhydrous (Sigma Aldrich, ≥98 % ), or

CaCl2 (Mallinckrodt, 74.0-78.0 % ACS grade). In experiments involving

112

Ethylenediaminetetraacetic acid (EDTA, Macron Fine Chemicals), EDTA was added after the desired concentration of inorganic salt had been added to the sample solution.

An ambient ocean water sample was collected for Inductively Coupled Plasma-Mass

Spectrometry and surface tension analysis (see below) from the coastal waters at La Jolla Shores

(La Jolla, CA). The water was collected into a 250 mL media bottle that had been cleaned in the same way as described above.

5.3.3 N2O5 and ClNO2 detection

N2O5 and ClNO2 were measured directly in the gas-phase using a Chemical Ionization

- Quadrupole Mass Spectrometer (CI-QMS) operating with I as the reagent ion. As such, N2O5

- -. was detected as the I adducts I N2O5, and detected at m/z 235 as described previously by Kercher

12 -. 35 at al.. Similarly, I ClNO2 was detected at m/z 208 for the Cl isotope.

Saturated NaCl solutions or 500mM NaCl solutions were used during each experiment as a reference calibration for ClNO2 yields. Previous work has shown that this method of

2 determining ClNO2 yields is consistent with previous methods used in the literature.

5.3.4 N2O5 generation and delivery

13 N2O5 was generated as described previously by Bertram et al.. Briefly, ozone is created via flowing zero air, combined with dry UHP N2, past a mercury UV light. This flow is then combined with an excess flow of NO2 in a mixing cell, where the gases are able to mix for approximately 2 minutes. Based on the ozone concentration titrated during the process, we estimate the N2O5 concentration as 75 ppb. The total combined flow of the products exiting the generation apparatus is 100 sccm. N2O5 was delivered to the solution surface using a method described previously.2 Briefly, A Teflon bottle cap was fitted with two Teflon Swagelok ports, allowing for the attachment of two Teflon tubes. The inlet port is bored through, allowing a 1/8"

Teflon tube delivering 100 sccm of N2O5, to reach within 2 cm of the solution surface. Given the

113 sample flow rate, surface area for reaction, and sample flow rate, we expect full reaction of all

N2O5 generated, and as such this experiment does not permit calculation of N2O5 reactivity, only

ClNO2 yields.

The 100 sccm flow exiting the reaction vessel is delivered via a 1/4 " Teflon tube, protected from light with aluminum foil, from the bottle headspace to the inlet of the CI-QMS. At the inlet of the instrument, the sample flow is combined with wet N2, generated by flowing dry N2 through a water bubbler, to keep the RH of the flow into the CI-QMS above 50% RH, so as to

- . prevent the detection of N2O5 and ClNO2 from being dependent on the I to I (H2O)n ratio in the instrument.

5.3.5 Ion Chromatography Analysis

Ion Chromatography (IC) analysis was performed on solutions of artificial reef salt,

MgCl2 anhydrous, MgCl2.6H2O (Macron) to assess the concentration magnesium within each sample.

5.3.6 Solution pH and Surface Tension Measurements

The pH of each magnesium chloride solution was determined by preparing fresh solutions and transferring an aliquot to a vial, where the pH was measured using an IQ Scientific

Instruments pH/mV/thermometer.

. Surface tension was measured for MgCl2 6H2O from both Sigma Aldrich and Macron,

MgCl2 anhydrous (Sigma Aldrich), coastal ocean water, and artificial reef salt solution.

Measurements were taken in duplicate with a Kruss K11 Tensiometer utilizing the Wilhelmy Pt plate method. The glass sampling cup and Wilhelmy plate were cleaned with ethanol and deionized water prior to measurement.

114

5.3.7 Inductively Coupled Plasma-Mass Spectrometry (ICP-MS) Analysis

Stock solutions of 3.5 ppt salinity artificial reef salt, 53 mM Mg2+ solutions made from

MgCl2 anhydrous, MgCl2.6H2O (Macron) and MgCl2.6H2O (Sigma Aldrich) separately, and a coastal ocean water sample as described above. This analysis was performed to assess the composition of each solution and to identify any possible differences in relative abundance of minor constituents.

10 L of each of the five solutions was transferred into a conical vial in a metal-free clean laboratory environment, and diluted with 10 mL 2% HNO3 solution. The diluted samples were analyzed using an inductively coupled plasma mass spectrometer (iCAPq quadropole ICP-

MS, Thermo Scientific) in duplicate. 2% HNO3 solution standard samples were run between each sample to ensure background levels remained constant. After collection, all HNO3 spectra were averaged together, and the average spectrum was subtracted from each sample spectrum to remove interferences from the dilution matrix.

5.4 Results and Discussion

The ClNO2 yields derived from N2O5 reaction with the ambient seawater samples, previously presented in Chapter 4, are shown again in Figure 5.1 (grey markers) to illustrate the typical ClNO2 yield values obtained for natural seawater (0.16-0.30). These seawater samples contain natural organics, however, upon flowing N2O5 over the surface artificial reef salt, a salt matrix that has no biological activity and should have minimal to no organic content, it shows a similar suppression in (ClNO2) down to 0.33 ± 0.04. This indicates the possibility of an inorganic mechanism that may contribute to the (ClNO2) suppression.

115

2 Figure 5.1 Yield of ClNO2 as a function of chlorophyll for real ocean water samples (grey) and for artificial reef salt mix solution (red). *Data from Chapter 4, Ryder et al. 2014, under review2

To investigate the role of major inorganic ions in regulating (ClNO2), a series of experiments were performed, where 500 mM NaCl solutions were doped with prevalent cationic components up to and exceeding their ambient concentrations. In one experiment magnesium, in the form of MgCl2 anhydrous was added, and in another experiment calcium was added in the form of CaCl2. The results of these experiments are shown in Figure 5.2, where the solid lines indicate the average ocean concentration of Ca2+ (light blue) and Mg2+ (light red). These results show that there is no reduction following the addition of CaCl2. However, upon the addition of

2+ MgCl2 anhydrous, the yield decreases down to 0.26 ± 0.1 at [Mg ] = 63 mM, after which the rate of decrease in (ClNO2) slows, and the yield values are within error of one another. This suggests a mechanism of suppression specific to Mg2+ presence in solution.

116

Figure 5.2 ClNO2 yield following MgCl2 anhydrous addition to a 500 mM NaCl solution (circles, panel A). 2+ The solid line indicates the typical concentration of Mg in the ocean. Panel B shows the ClNO2 yield following EDTA addition to a solution of 53 mM Mg2+ in 500 mM NaCl solution.

In order to ensure the result seen in the case of Mg2+ was a result of the metal cation and not large concentrations of organic material present as contamination in the salt, EDTA was added to a 53 mM MgCl2 anhydrous salt solution. Figure 5.3 shows the results of this addition as well as EDTA addition to artificial reef salt. In both cases, (ClNO2) increases with incremental addition of EDTA. In the case of MgCl2 anhydrous, the yield recovers up to 0.73 after 3.3 molar equivalents of EDTA are added while the artificial reef salt reaches a maximum of 0.51 after 4.5

EDTA equivalents. The lower recovery of the reef salt is likely a result of the composition of ions present in the artificial reef salt which can also complex with EDTA. One would expect a priori

117 that only one molar equivalent of EDTA would be required to chelate the Mg2+, however EDTA forms metal complexes most efficiently when in EDTA4- form, which occurs around pH =11. As seen in Table 5.1, the pH of MgCl2 anhydrous and artificial reef salt solutions are 6.4 and 7.8, respectively, indicating that EDTA is likely in a protonated form in these solutions, and not as effective in binding to the metal ions in solution. Despite the addition of more than 1 equivalent of EDTA to solution to reach peak recovery in each of these cases, it is apparent from this analysis that a metal cation is involved in the mechanism for (ClNO2) suppression.

Figure 5.3 ClNO2 yield as a function of EDTA molar equivalent additions to 53 mM MgCl2 anhydrous (blue) and artificial reef salt (green). The dashed lines are fits to serve as a guide.

118

Table 5.1 Sample solutions and their associated measured pH, [Mg2+] as measured by IC, and surface tension.

Solution pH in 500 mM [Mg2+] (mM) Surface Tension (mN/m)

NaCl

MgCl2 anhydrous 6.4 ± 0.2 62.6 ± 0.5 72.6 ± 0.14

MgCl2.6H2O (Macron) 6.0 ± 0.1 71.3 ± 0.5 72.7 ± 0

MgCl2.6H2O (Sigma Aldrich) N/A N/A 72.7 ± 0.42

Artificial reef salt 7.8 74.9 ± 0.6 73.3 ± 0.78

Coastal Ocean Water N/A N/A 73.4 ± 0.14

2+ To further explore this, we determined (ClNO2) for solutions of Mg originating from

. . MgCl2 6H2O, as it has been noted previously that MgCl2 6H2O are inherently cleaner than the anhydrous forms. The comparison of the two samples is shown in Figure 5.4, where, (ClNO2) is plotted as a function of MgCl2.6H2O (Macron) added to a 500 mM NaCl solution (grey). In the case of MgCl2.6H2O (Macron), no suppression in the yield of ClNO2 is observed at any point in the experiment. This was repeated with MgCl2.6H2O (Sigma Aldrich), and the same result was obtained (data not shown). To gain insight into differences between the anhydrous, hexahydrate, artificial reef salt and real ocean water, a number of analyses were performed as discussed below.

119

2+ Figure 5.4 ClNO2 yield as a function of [Mg ] added to a 500 mM NaCl solution for both MgCl2.6H2O (grey squares) and MgCl2 anhydrous (red squares). The oceanic concentration of magnesium is denoted by the blue line at 53 mM Mg2+.

2+ 2+. Once in solution, Mg should be found as Mg 6H2O independent of the initial form of

2+ magnesium that was added. To confirm the free Mg concentrations in each sample, MgCl2

. anhydrous, MgCl2 6H2O (Macron) and artificial reef salt solutions were analyzed by IC, the results of which are shown in Table 5.1. Although the absolute [Mg2+] for the three samples

2+ deviate by as much as 21 mM from the average concentration of Mg in the ocean, MgCl2 anhydrous and artificial reef salt exhibit the largest [Mg2+] difference from one another. Since

2+ these two salt solutions result in the same level of (ClNO2) suppression, the [Mg ] is likely not the driving force behind the lack of suppression seen in the MgCl2.6H2O (Macron) solution.

120

Both experimental and theoretical work has shown Mg2+ in solution has a low propensity toward the air/surface interface.14,10 Additionally, as was shown in Chapter 4,2 Cl- in the surface layers of solution also exists in diminished concentration. However, magnesium, due to its size and high charge density, does not form ion pairs in aqueous solution with Cl-, and thus cannot impact the distribution of chloride at the interface directly.15 Recent work has shown that the presence of magnesium in solution will influence the microsolvation of chloride present at the surface, which could impact its ability to interact during heterogeneous reactions.10 However, it is not clear at this stage how this would impact the observed differences in (ClNO2) for the two magnesium solutions.

Minofar and coworkers have also shown that small organics such acetate will reside at the interface, and dependent on their concentration, will pull Mg2+ up towards the surface.14 It might be possible that if soluble anionic organics are also present in solution, the attraction of

Mg2+ towards the hydrophobic organics at the surface also causes a net increase in the more soluble organics towards the surface layers, where Mg2+ resides. If this is the case, it is possible that the presence of Mg2+ in solution where organics are also present causes a net increase in

+ - organic concentration at the surface, possibly inhibiting NO2 reaction with Cl . If the MgCl2 anhydrous salt, due to being less pure, has more organic contaminants, that this could be causing the reduction in (ClNO2) observed.

To test for relative quantities of organics in each of the salt mixes, and to compare to a real system, surface tension measurements were performed on the MgCl2 salt samples, artificial reef salt, and a coastal water sample. The results of these measurements are shown in Table 5.1.

The coastal ocean water sample showed the highest average surface tension value at 73.4 ± 0.14, with the artificial reef salt producing the same value within error (73.3 ± 0.78). The magnesium salts showed lower surface tension values (72.6 ± 0.14 to 72.7 ± 0.42), which could be due to two possible effects. Either the lower salinity of the MgCl2 solutions as compared to the artificial reef

121 salt and coastal ocean sample, is resulting in surface tension values that are closer to deionized water (71.99 at 25 C)16 than to the higher salinity salt samples, or organics are present in these compounds, reducing the surface tension of the solutions. We would expect the former as a more likely explanation, however, since real ocean water contains organic content (average =70 M carbon), and the surface tension value is still higher than that of deionized water, more than surface tension analysis is needed to infer organic content of these salts. Despite this, we can conclude from the surface tension measurements that sufficiently high concentrations of reactive organic surfactants are likely not present in any of the salt solutions.

To probe for inorganic contaminants that might be di- or trivalent in nature and may exhibit preference to reside in the bulk over the surface and could draw chloride away from the interface, as was discussed earlier for Mg2+, ICP-MS analysis was performed. Specifically we were interested in trace metal contaminants present in real ocean water, artificial reef salt, and

MgCl2 anhydrous and less or not present in the hexahydrate MgCl2 salts. Figure 5.5 shows the

ICP-MS data as a bar plot. Here we confirm that magnesium is present in comparable levels across all samples, in agreement with the previously discussed IC analysis (Figure 5.5A-E).

Figure 5.6 shows a closer view of ions present in lower concentrations and their relative abundances in each of the salt samples. Although there are variations between the contents of the various salts, two main points can be concluded from this data set. Firstly, both the anhydrous and hexahydrate forms of MgCl2 contain trace contaminants. Secondly, over the duplicate runs, the ion that was reproducibly present in the ocean, artificial reef salt, and MgCl2 anhydrous and less so in the hexahydrate salts was 57Fe (it should be noted that the 54Fe signal is not included due to a known interference at that m/z), shown in Figure 5.6F. However, it is evident that the

57 concentration of Fe is lower in the case of MgCl2 anhydrous versus in the real ocean and

57 artificial reef salt samples. MgCl2.6H2O (Macron) also appears to contain a small amount of Fe,

122 however, it is not immediately clear if this is real or carryover/a shoulder from a larger adjacent peak.

Figure 5.5 ICP-MS analysis of coastal ocean water (A), and artificial reef salt (B), MgCl2.6H2O, Sigma Aldrich (C), MgCl2.6H2O, Macron (D), MgCl2 anhydrous (E).

If this is indeed an ion that is absent in the hexahydrate samples, it is possible that this ion is capable of drawing Cl- away from the interface, whereas Mg2+ is not (see above). No matter what

123 the cause of the suppression from the salt solutions, even if due to the presence of a contaminant, the fact that a reduction of (ClNO2) down to the same level as in the ocean strongly suggests the same mechanistic suppression is responsible.

Figure 5.6 A closer view of Figure 5.5, showing ions present in lower concentrations. Inset Panel F shows a closer view of the 57Fe peak signal.

124

5.5 Future work/ challenges

The results of this work are ongoing and although remain inconclusive, indicate a possible enhancement of organics at the interface due to the presence of divalent cations.

Questions that still remain to be answered include: 1) What is the exact mechanism through

2+ which Mg is influencing the ClNO2 production? 2) Is the solvation shell of the hexahydrate salt preventing Mg2+ from affecting the microsolvation environment of Cl- ? 3) Is another di- or trivalent ion impacting Cl- availability in the first 50nm of the interface? A main step forward in answering these questions is to first determine conclusively why the anhydrous and hexahydrate forms of the salts exhibit different behavior and whether it is due to the magnesium form, or whether it is due to a trace contaminant effect. One way to elucidate this is to hydrate the MgCl2 anhydrous solution, re-crystallize the salt, and use this purified form of the slat in the same manner as before. To test the effect of iron on (ClNO2) suppression, an experiment similar to that described for MgCl2 should be repeated with FeCl3.

5.6 Acknowledgments

Chapter 5, in part, is currently being prepared for submission of the material.

Olivia S. Ryder and Timothy H. Bertram. The dissertation author was the primary investigator and author of this material.

This work was supported by the National Science Foundation via the Center for Aerosol

Impact on Climate and the Environment, a Center for Chemical Innovation (NSF CHE1305427).

The authors would like to gratefully thank Scott Wilson and Dr. Skip Pomeroy for IC analysis,

Dr. James Day for ICP-MS analysis, Dr. Dale Stokes and Dr. Grant Deane for use of the surface tensiometer used in this work, Dr. Kim Prather, Dr. Michael Tauber and Dr. Andreae Meinrat for helpful discussions. O.S.R gratefully acknowledges a Graduate Research Fellowship from the

National Science Foundation (2011-2014).

125

5.7 References

(1) Kim, M. J.; Farmer, D. K.; Bertram, T. H. A Controlling Role for the Air−sea Interface in the Chemical Processing of Reactive Nitrogen in the Coastal Marine Boundary Layer. Proc. Natl. Acad. Sci. 2014, 111 (11), 3943–3948.

(2) Ryder, O. S.; Campbell, Nicole R.; Shaloski, Michael; Al-Mashat, Hashim; Nathanson, Gilbert M.; Bertram, Timothy H. On the Role of Organics in Regulating ClNO2 Production at the Air-Sea Interface. J. Phys. Chem. A 2014, in review.

(3) Ogawa, H.; Tanoue, E. Dissolved Organic Matter in Oceanic Waters. J. Oceanogr. 2003, 59 (2), 129–147.

(4) Hansell, D. A. Chapter 15 - DOC in the Global Ocean Carbon Cycle. In of Marine Dissolved Organic Matter; Carlson, D. A. H. A., Ed.; Academic Press: San Diego, 2002; pp 685 – X.

(5) Millero, F. J.; Feistel, R.; Wright, D. G.; McDougall, T. J. The Composition of Standard Seawater and the Definition of the Reference-Composition Salinity Scale. Deep Sea Res. Part Oceanogr. Res. Pap. 2008, 55 (1), 50–72.

(6) Marcet, A. Abstract of Experiments on the Specific Gravity, Temperature, and Saline Contents of Sea-Water, &c.; Printed for A. Constable: Edinburgh, 1820.

(7) Curry, J. A.; Webster, P. J. Thermodynamics of Atmospheres and Oceans; Academic Press, 1999. (8) Jungwirth, P.; Tobias, D. J. Specific Ion Effects at the Air/Water Interface. Chem. Rev. 2006, 106 (4), 1259–1281.

(9) Jungwirth, P.; Tobias, D. J. Molecular Structure of Salt Solutions: A New View of the Interface with Implications for Heterogeneous Atmospheric Chemistry. J. Phys. Chem. B 2001, 105 (43), 10468–10472.

(10) Callahan, K. M.; Casillas-Ituarte, N. N.; Xu, M.; Roeselová, M.; Allen, H. C.; Tobias, D. J. Effect of Magnesium Cation on the Interfacial Properties of Aqueous Salt Solutions. J. Phys. Chem. A 2010, 114 (32), 8359–8368.

(11) Hua, W.; Verreault, D.; Huang, Z.; Adams, E. M.; Allen, H. C. Cation Effects on Interfacial Water Organization of Aqueous Chloride Solutions. I. Monovalent Cations: Li+, Na+, K+, and NH4+. J. Phys. Chem. B 2014, 118 (28), 8433–8440.

(12) Kercher, J. P.; Riedel, T. P.; Thornton, J. A. Chlorine Activation by N2O5: Simultaneous, in Situ Detection of ClNO2 and N2O5 by Chemical Ionization Mass Spectrometry. Atmos Meas Tech 2009, 2 (1), 193–204.

(13) Bertram, T. H.; Thornton, J. A.; Riedel, T. P. An Experimental Technique for the Direct Measurement of N2O5 Reactivity on Ambient Particles. Atmos Meas Tech 2009, 2 (1), 231–242.

126

(14) Minofar, B.; Vácha, R.; Wahab, A.; Mahiuddin, S.; Kunz, W.; Jungwirth, P. Propensity for the Air/Water Interface and Ion Pairing in Magnesium Acetate vs Magnesium Nitrate Solutions: Molecular Dynamics Simulations and Surface Tension Measurements. J. Phys. Chem. B 2006, 110 (32), 15939–15944.

(15) Casillas-Ituarte, N. N.; Callahan, K. M.; Tang, C. Y.; Chen, X.; Roeselová, M.; Tobias, D. J.; Allen, H. C. Surface Organization of Aqueous MgCl2 and Application to Atmospheric Marine Aerosol Chemistry. Proc. Natl. Acad. Sci. 2010, 107 (15), 6616–6621.

(16) Vargaftik, N. B.; Volkov, B. N.; Voljak, L. D. International Tables of the Surface Tension of Water. J. Phys. Chem. Ref. Data 1983, 12 (3), 817–820.