ArticlePDF Available

Assessing the impact of anthropogenic pollution on isoprene-derived secondary organic aerosol formation in PM2.5 collected from the Birmingham, Alabama ground site during the 2013 Southern Oxidant and Aerosol Study

Authors:

Abstract

In the southeastern U.S., substantial emissions of isoprene from deciduous trees undergo atmospheric oxidation to form secondary organic aerosol (SOA) that contributes to fine particulate matter (PM2.5). Laboratory studies have revealed that anthropogenic pollutants, such as sulfur dioxide (SO2), oxides of nitrogen (NOx), and aerosol acidity, can enhance SOA formation from the hydroxyl radical (OH)-initiated oxidation of isoprene; however, the mechanisms by which specific pollutants enhance isoprene SOA in ambient PM2.5 remain unclear. As one aspect of an investigation to examine how anthropogenic pollutants influence isoprene-derived SOA formation, high-volume PM2.5 filter samples were collected at the Birmingham, Alabama (BHM) ground site during the 2013 Southern Oxidant and Aerosol Study (SOAS). Sample extracts were analyzed by gas chromatography/electron ionization-mass spectrometry (GC/EI-MS) with prior trimethylsilylation and ultra performance liquid chromatography coupled to an electrospray ionization high-resolution quadrupole time-of-flight mass spectrometry (UPLC/ESI-HR QTOFMS) to identify known isoprene SOA tracers. Tracers quantified using both surrogate and authentic standards were compared with collocated gas- and particle-phase data as well as meteorological data provided by the Southeastern Aerosol Research and Characterization (SEARCH) network to assess the impact of anthropogenic pollution on isoprene-derived SOA formation. Results of this study reveal that isoprene-derived SOA tracers contribute a substantial mass fraction of organic matter (OM) (~7 to ~20%). Isoprene-derived SOA tracers correlated with sulfate (SO42-) (r2 = 0.34, n = 117), but not with NOx. Moderate correlation between methacrylic acid epoxide and hydroxymethyl-methyl-α-lactone (MAE/HMML)-derived SOA tracers and nitrate radical production (P[NO3]) (r2 = 0.57, n = 40) were observed during nighttime, suggesting a potential role of NO3 radical in forming this SOA type. However, the nighttime correlation of these tracers with nitrogen dioxide (NO2) (r2 = 0.26, n = 40) was weaker. Ozone (O3) correlated strongly with MAE/HMML-derived tracers (r2 = 0.72, n = 30) and moderately with 2-methyltetrols (r2 = 0.34, n = 15) during daytime only, suggesting that a fraction of SOA formation could occur from isoprene ozonolysis in urban areas. No correlation was observed between aerosol pH and isoprene-derived SOA. Lack of correlation between aerosol acidity and isoprene-derived SOA indicates that acidity is not a limiting factor for isoprene SOA formation at the BHM site as aerosols were acidic enough to promote multiphase chemistry of isoprene-derived epoxides throughout the duration of the study. All in all, these results confirm the reports that anthropogenic pollutants enhance isoprene-derived SOA formation.
1
Assessing the impact of anthropogenic pollution on isoprene-derived secondary organic 1
aerosol formation in PM2.5 collected from the Birmingham, Alabama ground site during the 2
2013 Southern Oxidant and Aerosol Study 3
4
W. Rattanavaraha1, K. Chu1, S. H. Budisulistiorini1,a, M. Riva1, Y.-H. Lin1,b, E. S. Edgerton2, K. 5
Baumann2, S. L. Shaw3, H. Guo4, L. King4, R. J. Weber4, E. A. Stone5, M. E. Neff5, J. H. 6
Offenberg6, Z. Zhang1, A. Gold1, and J. D. Surratt1,* 7
8
1 Department of Environmental Sciences and Engineering, Gillings School of Global Public 9
Health, The University of North Carolina at Chapel Hill, Chapel Hill, NC, USA 10
2 Atmospheric Research & Analysis, Inc., Cary, NC, USA 11
3 Electric Power Research Institute, Palo Alto, CA, USA 12
4 Earth and Atmospheric Science, Georgia Institute of Technology, Atlanta, GA, USA 13
5 Department of Chemistry, University of Iowa, Iowa City, IA, USA 14
6 Human Exposure and Atmospheric Sciences Division, United States Environmental Protection 15
Agency, Research Triangle Park, NC, USA 16
a now at: Earth Observatory of Singapore, Nanyang Technological University, Singapore 17
b now at: Michigan Society of Fellows, Department of Chemistry, University of Michigan, Ann 18
Arbor, MI, USA 19
20
* To whom correspondence should be addressed. Email: surratt@unc.edu 21
For Submission to: Atmospheric Chemistry & Physics Discussions 22
23
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
2
Abstract 24
In the southeastern U.S., substantial emissions of isoprene from deciduous trees undergo 25
atmospheric oxidation to form secondary organic aerosol (SOA) that contributes to fine particulate 26
matter (PM2.5). Laboratory studies have revealed that anthropogenic pollutants, such as sulfur 27
dioxide (SO2), oxides of nitrogen (NOx), and aerosol acidity, can enhance SOA formation from 28
the hydroxyl radical (OH)-initiated oxidation of isoprene; however, the mechanisms by which 29
specific pollutants enhance isoprene SOA in ambient PM2.5 remain unclear. As one aspect of an 30
investigation to examine how anthropogenic pollutants influence isoprene-derived SOA 31
formation, high-volume PM2.5 filter samples were collected at the Birmingham, Alabama (BHM) 32
ground site during the 2013 Southern Oxidant and Aerosol Study (SOAS). Sample extracts were 33
analyzed by gas chromatography/electron ionization-mass spectrometry (GC/EI-MS) with prior 34
trimethylsilylation and ultra performance liquid chromatography coupled to an electrospray 35
ionization high-resolution quadrupole time-of-flight mass spectrometry (UPLC/ESI-HR-36
QTOFMS) to identify known isoprene SOA tracers. Tracers quantified using both surrogate and 37
authentic standards were compared with collocated gas- and particle-phase data as well as 38
meteorological data provided by the Southeastern Aerosol Research and Characterization 39
(SEARCH) network to assess the impact of anthropogenic pollution on isoprene-derived SOA 40
formation. Results of this study reveal that isoprene-derived SOA tracers contribute a substantial 41
mass fraction of organic matter (OM) (~7 to ~20%). Isoprene-derived SOA tracers correlated with 42
sulfate (SO42-) (r2 = 0.34, n = 117), but not with NOx. Moderate correlation between methacrylic 43
acid epoxide and hydroxymethyl-methyl-α-lactone (MAE/HMML)-derived SOA tracers and 44
nitrate radical production (P[NO3]) (r2 = 0.57, n = 40) were observed during nighttime, suggesting 45
a potential role of NO3 radical in forming this SOA type. However, the nighttime correlation of 46
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
3
these tracers with nitrogen dioxide (NO2) (r2 = 0.26, n = 40) was weaker. Ozone (O3) correlated 47
strongly with MAE/HMML-derived tracers (r2 = 0.72, n = 30) and moderately with 2-methyltetrols 48
(r2 = 0.34, n = 15) during daytime only, suggesting that a fraction of SOA formation could occur 49
from isoprene ozonolysis in urban areas. No correlation was observed between aerosol pH and 50
isoprene-derived SOA. Lack of correlation between aerosol acidity and isoprene-derived SOA 51
indicates that acidity is not a limiting factor for isoprene SOA formation at the BHM site as 52
aerosols were acidic enough to promote multiphase chemistry of isoprene-derived epoxides 53
throughout the duration of the study. All in all, these results confirm the reports that anthropogenic 54
pollutants enhance isoprene-derived SOA formation. 55
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
4
1. Introduction 56
Fine particulate matter, suspensions of liquid or solid aerosol in a gaseous medium that are 57
less than or equal to 2.5 μm in diameter (PM2.5), play a key role in physical and chemical 58
atmospheric processes. They influence climate patterns both directly, through the absorption and 59
scattering of solar and terrestrial radiation, and indirectly, through cloud formation (Kanakidou et 60
al., 2005). In addition to climatic effects, PM2.5 has been demonstrated to pose a potential human 61
health risk through inhalation exposure (Pope and Dockery, 2006; Hallquist et al., 2009). Despite 62
the strong association of PM2.5 with climate change and environmental health, there remains a need 63
to more fully resolve its composition, sources, and chemical formation processes in order to 64
develop effective control strategies to address potential hazards in a cost-effective manner 65
(Hallquist et al., 2009; Boucher et al., 2013; Nozière et al., 2015). 66
Atmospheric PM2.5 are comprised in a large part (up to 90% by mass in some locations), 67
of organic matter (OM) (Carlton et al., 2009; Hallquist et al., 2009). OM can be derived from many 68
sources. Primary organic aerosol (POA) is emitted from both natural (e.g., fungal spores, 69
vegetation, vegetative detritus) and anthropogenic sources (fossil fuel and biomass burning) prior 70
to atmospheric processing. As a result of large anthropogenic sources, POA is abundant largely in 71
urban areas. Processes such as biomass burning and combustion also yield volatile organic 72
compounds (VOCs), which have high vapor pressures and can undergo atmospheric oxidation to 73
form secondary organic aerosol (SOA) through gas-to-particle phase partitioning (condensation or 74
nucleation) with subsequent particle-phase (multiphase) chemical reactions (Grieshop et al., 75
2009). 76
At around 600 Tg emitted per year into the atmosphere, isoprene (2-methyl-1,3-butadiene, 77
C5H8) is the most abundant volatile non-methane hydrocarbon (Guenther et al., 2012). The 78
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
5
abundance of isoprene is particularly high in the southeastern U.S. due to emissions from broadleaf 79
deciduous tree species (Guenther et al., 2006). Research over the last decade has revealed that 80
isoprene, via hydroxyl radical (OH)-initiated oxidation, is a major source of SOA (Claeys et al., 81
2004; Edney et al., 2005; Kroll et al., 2005 ; Kroll et al., 2006; Surratt et al., 2006; Lin et al., 2012; 82
Lin et al., 2013a). In addition, it is known that SOA formation is enhanced by anthropogenic 83
emissions, namely oxides of nitrogen (NOx) and sulfur dioxide (SO2), that are a source of acidic 84
aerosol onto which photochemical oxidation products of isoprene are reactively taken up to yield 85
a variety of SOA products (Edney et al., 2005; Kroll et al., 2006; Surratt et al., 2006; Surratt et al., 86
2007b; Surratt et al., 2010; Lin et al., 2013b;) . 87
Recent work has begun to elucidate some of the critical intermediates of isoprene oxidation 88
that lead to SOA formation through acid-catalyzed heterogeneous chemistry (Kroll et al., 2005; 89
Surratt et al., 2006). Under low-NOx conditions, such as in a pristine environment, isomeric 90
isoprene epoxydiols (IEPOX) have been demonstrated to be critical to the formation of isoprene 91
SOA. On advection of IEPOX to an urban environment and mixing with anthropogenic emissions 92
of acidic sulfate aerosol, SOA formation is enhanced (Surratt et al., 2006; Lin et al., 2012; Lin et 93
al., 2013b). This pathway has been shown to yield 2-methyltetrols as major SOA constituents of 94
ambient PM2.5 (Claeys et al, 2004; Surratt et al., 2010; Lin et al., 2012). Further work has revealed 95
a number of additional IEPOX-derived SOA tracers, including C5-alkene triols (Wang et al., 2005; 96
Lin et al., 2012), cis- and trans-3-methyltetrahydrofuran-3,4-diols (3-MeTHF-3,4-diols) (Lin et 97
al., 2012; Zhang et al., 2012), IEPOX-derived organosulfates (OSs) (Lin et al., 2012), and IEPOX-98
derived oligomers (Lin et al., 2014). Some of the IEPOX-derived oligomers have been shown to 99
contribute to aerosol components known as brown carbon that absorb light in the near ultraviolet 100
(UV) and visible ranges (Lin et al., 2014). Under high-NOx conditions, such as encountered in an 101
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
6
urban environment, isoprene is oxidized to methacrolein and SOA formation occurs via the further 102
oxidation of methacrolein (MACR) (Kroll et al., 2006; Surratt et al., 2006) to methacryloyl 103
peroxynitrate (MPAN) (Chan et al., 2010; Surratt et al., 2010; Nguyen et al., 2015). It has recently 104
been shown that when MPAN is oxidized by OH it yields at least two SOA precursors, methacrylic 105
acid epoxide (MAE) and hydroxymethyl-methyl-α-lactone (HMML) (Surratt et al., 2006; Surratt 106
et al., 2010; Lin et al., 2013a; Nguyen et al., 2015). Whether SOA precursors are formed under 107
high- or low-NOx conditions, aerosol acidity is a critical parameter that enhances the reaction 108
kinetics through acid-catalyzed reactive uptake and multiphase chemistry of either IEPOX or 109
MAE/HMML ( Surratt et al., 2007b; Surratt et al., 2010; Lin et al., 2013b). 110
Due to the considerable emissions of isoprene, an SOA yield of even 1% would contribute 111
significantly to ambient SOA (Carlton et al., 2009; Henze et al., 2009). This conclusion is 112
supported by measurements showing that up to a third of total fine OA mass can be attributed to 113
IEPOX-derived SOA tracers in Atlanta, GA (JST) during summer months (Budisulistiorini et al., 114
2013; Budisulistiorini et al., 2015). A recent study in Yorkville, GA (YRK), similarly found that 115
IEPOX-derived SOA tracers comprised 12-19% of the fine OA mass (Lin et al., 2013b). Another 116
SOAS site at Centreville, Alabama (CTR) revealed IEPOX-SOA contributed 18% of total OA 117
mass (Xu et al., 2015). The individual ground sites corroborate recent aircraft-based measurements 118
made in the Studies of Emissions and Atmospheric Composition, Clouds, and Climate Coupling 119
by Regional Surveys (SEAC4RS) aircraft campaign, which estimates an IEPOX-SOA contribution 120
of 32% to OA mass in the southeastern U.S. (Hu et al., 2015). 121
It is clear from the field studies discussed above that particle-phase chemistry of isoprene-122
derived oxidation products plays a large role in atmospheric SOA formation. However, much 123
remains unknown regarding the exact nature of its formation, limiting the ability of models to 124
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
7
accurately account for isoprene SOA (Carlton et al., 2010b; Foley et al., 2010). Currently, 125
traditional air quality models in the southeastern U.S. do not incorporate detailed particle-phase 126
chemistry of isoprene oxidation products (IEPOX or MAE/HMML) and generally under-predict 127
isoprene SOA formation (Carlton et al., 2010a). Recent work demonstrates that incorporating the 128
specific chemistry of isoprene epoxide precursors into models increases the accuracy of isoprene 129
SOA prediction (Pye et al., 2013; Karambelas et al., 2014), suggesting that understanding the 130
formation mechanisms of biogenic SOA, especially with regard to the effects of anthropogenic 131
emissions, such as NOx and SO2, will be key to more accurate models. More accurate models are 132
needed in order to devise cost-effective control strategies for reducing PM2.5 levels. Since isoprene 133
is primarily biogenic in origin, and therefore not controllable, the key to understanding the public 134
health and environmental implications of isoprene SOA lies in resolving the effects of 135
anthropogenic pollutants. 136
This study presents results from the 2013 Southeastern Oxidant and Aerosol Study 137
(SOAS), where several well-instrumented ground sites dispersed throughout the southeastern U.S. 138
made intensive gas- and particle-phase measurements from June 1 – July 16, 2013. The primary 139
purpose of this campaign was to examine, in greater detail, the formation mechanisms, 140
composition, and properties of biogenic SOA, including the effects of anthropogenic emissions. 141
This study pertains specifically to the results from the BHM ground site, where the city’s ample 142
urban emissions mix with biogenic emissions from the surrounding rural areas, creating an ideal 143
location to investigate such interactions. The results presented here focus on analysis of PM2.5 144
collected on filters during the campaign by gas chromatography interfaced to electron ionization-145
mass spectrometry (GC/EI-MS) and ultra performance liquid chromatography interfaced with 146
electrospray ionization high-resolution quadrupole time-of-flight mass spectrometry (UPLC/ESI-147
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
8
HR-QTOFMS). The analysis of PM2.5 was conducted in order to measure quantities of known 148
isoprene SOA tracers and using collocated air quality and meteorological measurements to 149
investigate how anthropogenic pollutants including NOx, SO2, aerosol acidity (pH), PM2.5 sulfate 150
(SO42-), and O3 affect isoprene SOA formation. These results, along with the results presented 151
from similar studies during the 2013 SOAS campaign, seek to elucidate the chemical relationships 152
between anthropogenic emissions and isoprene SOA formation in order to provide better 153
parameterizations needed to improve the accuracy of air quality models in this region of the U.S. 154
2. Methods 155
2.1. Site description and collocated data 156
Filter samples were collected in the summer of 2013 as part of the SOAS field campaign 157
at the BHM ground site (33.553N, 86.815W). In addition to the SOAS campaign, the site is also 158
part of the Southeastern Aerosol Research and Characterization Study (SEARCH) (Figure S1 of 159
the Supplement), an observation and monitoring program initiated in 1998. SEARCH and this site 160
are described elsewhere in detail (Hansen et al., 2003; Edgerton et al., 2006). The BHM site is 161
surrounded by significant transportation and industrial sources of PM. West of BHM are US-31 162
and I-65 highways. To the north, northeast and southwest of BHM several coking ovens and an 163
iron pipe foundry are located (Hansen et al., 2003). 164
2.2. High-Volume filter sampling and analysis methods 165
2.2.1. High-Volume filter sampling 166
From June 1 July 16, 2013, PM2.5 samples were collected onto TissuquartzTM Filters 167
(8 x 10 in, Pall Life Sciences) using high-volume PM2.5 samplers (Tisch Environmental) operated 168
at 1 m3 min-1 at ambient temperature described in detail elsewhere (Budisulistiorini et al. 2015; 169
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
9
Riva et al., 2015). All quartz filters were pre-baked prior to collection. The procedure consisted of 170
baking filters at 550 °C for 18 hours followed by cooling to 25 °C over 12 hours. 171
The sampling schedule is given in Table 1. Either two or four samples were collected per 172
day. The regular schedule consisted of two samples per day, one during the day, the second at 173
night, each collected for 11 hours. On intensive sampling days, four samples were collected, with 174
the single daytime sample being subdivided into three separate periods. The intensive sampling 175
schedule was conducted on days when high levels of isoprene, SO42- and NOx where forecast by 176
the National Center for Atmospheric Research (NCAR) using the Flexible Particle dispersion 177
model (FLEXPART) (Stohl et al., 2005) and Model for Ozone and Related Chemical Tracers 178
(MOZART) (Emmons et al., 2010) simulations. Details of these simulations have been 179
summarized in Budisulistiorini et al. (2015); however, these model data were only used 180
qualitatively to determine the sampling schedule. The intensive collection frequency allowed 181
enhanced time resolution for offline analysis to examine the effect of anthropogenic emissions on 182
the evolution of isoprene SOA tracers throughout the day. 183
In total, 120 samples were collected throughout the field campaign with a field blank filter 184
collected every 10 days to identify errors or contamination in sample collection and analysis. All 185
filters were stored at -20 °C in the dark until extraction and analysis. In addition to filter sampling 186
of PM2.5, SEARCH provided a suite of additional instruments at the site collecting measurements 187
of a variety of variables, including meteorology, gas, and continuous PM monitoring. The variables 188
with respective instrumentation are summarized in Table S1 of the Supplement. 189
2.2.2. Isoprene-derived SOA analysis by GC/EI-MS 190
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
10
SOA collected in the field on quartz filters was extracted and isoprene tracers quantified 191
by GC/EI-MS with prior trimethylsilylation. A 37-mm diameter circular punch from each filter 192
was extracted in a pre-cleaned scintillation vial with 20 mL of high-purity methanol (LCMS 193
CHROMASOLV-grade, Sigma-Aldrich) by sonication for 45 minutes. The extracts were filtered 194
through PTFE syringe filters (Pall Life Science, Acrodisc®, 0.2-µm pore size) to remove insoluble 195
particles and residual quartz fibers. The filtrate was then blown dry under a gentle stream of N2 at 196
room temperature. The dried residues were immediately trimethylsilylated by reaction with 100 197
μL of BSTFA + TMCS (99:1 v/v, Supelco) and 50 μL of pyridine (anhydrous, 99.8 %, Sigma-198
Aldrich) at 70 °C for 1 hour. Derivatized samples were analyzed within 24 hours after 199
trimethylsilylation using a Hewlett-Packard (HP) 5890 Series II Gas Chromatograph coupled to a 200
HP 5971A Mass Selective Detector. The gas chromatograph was equipped with an Econo-Cap®-201
EC®-5 Capillary Column (30 m x 0.25 mm i.d.; 0.25-μm film thickness) to separate trimethylsilyl 202
derivatives before MS detection. 1 μL aliquots were injected onto the column. Operating 203
conditions and procedures have been described elsewhere (Surratt et al., 2010). 204
Extraction efficiency was assessed and taken into account for the quantification of all SOA 205
tracers. Efficiency was determined by analyzing 4 pre-baked filters spiked with 50 ppmv of 2-206
methyltetrols, 2-methylglyceric acid, levoglucosan, and cis- and trans-3-MeTHF-3,4-diols. 207
Extraction efficiency was above 90% and used to correct the quantification of samples. Extracted 208
ion chromatograms (EICs) of m/z 262, 219, 231, 335 were used to quantify the cis-/trans-3-209
MeTHF-3,4-diols, 2-methyltetrols and 2-methylglyceric acid, C5-alkene triols, and IEPOX-210
dimers, respectively (Surratt et al., 2006). 211
2-Methyltetrols were quantified using an authentic reference standard that consisted of a 212
mixture of racemic diasteroisomers. Similarly, 3-MeTHF-3,4-diol isomers were also quantified 213
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
11
using authentic standards; however, 3-MeTHF-3,4-diol isomers were detected in few field 214
samples. 2-Methylglyceric acid was also quantified using an authentic standard. Procedures for 215
synthesis of the 2-methyltetrols, 3-MeTHF-3,4-diol isomers, and 2-methylglyceric acid have been 216
described elsewhere (Zhang et al., 2012; Budisulistiorini et al., 2015). C5-alkene triols and IEPOX-217
dimers were quantified using the average response factor of the 2-methyltetrols. 218
2.2.3. Isoprene-derived SOA analysis by UPLC/ ESI-HR-QTOFMS 219
A 37-mm diameter circular punch from each quartz filter was extracted following the same 220
procedure described in section 2.2.1 for GC/EI-MS analysis. The dried residues were reconstituted 221
with 150 µl of a 50:50 (v/v) solvent mixture of methanol (LC-MS CHROMASOVL-grade, Sigma-222
Aldrich) and high-purity water (Milli-Q, 18.2 MΩ). The extracts were immediately analyzed by 223
the UPLC/ESI-HR-QTOFMS (6520 Series, Agilent) operated in the negative ion mode. Detailed 224
operating conditions have been described elsewhere (Riva et al., 2015). Mass spectra were 225
acquired at a mass resolution 7000-8000 over the range m/z 200 – 400. 226
Extraction efficiency was determined by analyzing 3 pre-baked filters spiked with propyl 227
sulfate and octyl sulfate (electronic grade, City Chemical LLC). Extraction efficiencies were in the 228
range 86 95%. EICs of m/z 215, 333 and 199 were used to quantify the IEPOX-derived OS, 229
IEPOX-derived dimer OS and the MAE/HMML-derived OS, respectively (Surratt et al., 2007a). 230
EICs were generated with a ± 5 ppm tolerance. All accurate masses for all measured 231
organosulfates were within ± 5 ppm. For simplicity, only the nominal masses are reported in the 232
text when describing these products. IEPOX-derived OS and IEPOX-derived dimer OS were 233
quantified by authentic standards (Zhang et al., 2012). The MAE/HMML-derived OS was 234
quantified using authentic MAE/HMML-derived OS synthesized in-house by a procedure to be 235
described in a forthcoming publication (1H NMR trace, Figure S2). 236
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
12
EICs of of m/z 155, 169 and 139 were used to quantify the glyoxal-derived OS, 237
methylglyoxal-derived OS, and the hydroxyacetone-derived OS, respectively (Surratt et al., 238
2007a). In addition, EICs of m/z 211, 260 and 305 were used to quantify other known isoprene-239
derived OSs (Surratt et al., 2007a). Glycolic acid sulfate synthesized in-house was used as a 240
standard to quantify the glyoxal-derived OS (Galloway et al., 2009) and propyl sulfate, was used 241
as a surrogate standard to quantify the remaining isoprene-derived OSs. 242
2.2.4. OC and WSOC analysis 243
A 1.5 cm2 square punch from each quartz filter was analyzed for total organic carbon (OC) 244
and elemental carbon (EC) by the thermal-optical method (Birch and Cary, 1996) on a Sunset 245
Laboratory OC/EC instrument (Tigard, OR) at the National Exposure Research Laboratory 246
(NERL) at the U.S. Environmental Protection Agency, Research Triangle Park, NC. The details 247
of the instrument and analytical method have been described elsewhere (Birch and Cary, 1996). In 248
addition to the internal calibration using methane gas, four different mass concentrations of sucrose 249
solution were used to verify the accuracy of instrument during the analysis. 250
Water-soluble organic carbon (WSOC) was measured in aqueous extracts of quartz fiber filter 251
samples using a total organic carbon (TOC) analyzer (Sievers 5310C, GE Water & Power) 252
equipped with an inorganic carbon remover (Sievers 900). To maintain low background carbon 253
levels, all glassware used was washed with water, soaked in 10% nitric acid, and baked at 500˚C 254
for 5 h and 30 min prior to use. Samples were extracted in batches that consisted of 12-21 PM2.5
255
samples and field blanks, one laboratory blank, and one spiked solution. A 17.3 cm2 filter portion 256
was extracted with 15 mL of purified water (> 18 MΩ, Barnstead Easypure II, Thermo Scientific) 257
by ultra-sonication (Branson 5510). Extracts were then passed through a 0.45 µm PTFE filter to 258
remove insoluble particles. The TOC analyzer was calibrated using potassium hydrogen phthalate 259
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
13
(KHP, Sigma Aldrich) and was verified daily with sucrose (Sigma Aldrich). Samples and standards 260
were analyzed in triplicate; the reported values correspond to the average of the second and third 261
trials. Spiked solutions yielded recoveries that averaged (± one standard deviation) 96 ± 5 % (n = 262
9). All ambient concentrations were field blank subtracted. 263
2.2.5. Estimation of aerosol pH by ISORROPIA 264
Aerosol pH was estimated using a thermodynamic model, ISORROPIA-II (Nenes et al., 265
1998). SO42-, nitrate (NO3-), and ammonium (NH4+) ion concentrations measured in PM2.5
266
collected from BHM, as well as relative humidity (RH), temperature and gas-phase ammonia 267
(NH3) were used as inputs into the model. These variables were obtained from the SEARCH 268
network at BHM, which collected the data during the period covered by the SOAS campaign. The 269
ISORROPIA-II model estimates particle hydronium ion concentration per unit volume of air (H+, 270
μg m−3), aerosol liquid water content (LWC, μg m−3), and aqueous aerosol mass concentration (μg 271
m−3). The model-estimated parameters were used in the following formula to calculate the aerosol 272
pH: 273
Aerosol pH = −log= −log( 
 ×  × 1000 ) 274
where is H+ activity in the aqueous phase (mol L-1), LMASS is total liquid-phase aerosol mass 275
(μg m−3) and  is aerosol density. Details of the ISORROPIA-II model and its ability to predict 276
pH, LWC, and gas-to-particle partitioning are not the focus of this study and are discussed 277
elsewhere and (Fountoukis et al., 2009). 278
2.2.6. Estimation of nighttime NO3 279
Nitrate radical (NO3) production (P[NO3]) was calculated using the following equation: 280
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
14
[] = [][] 281
where [NO2] and [O3] correspond to the measured ambient NO2 and O3 concentrations (mol 282
cm-3), respectively, and k is the temperature-dependent rate constant (Herron and Huie, 1974; 283
Graham and Johnston, 1978). Since no direct measure of NO3 radical was made at this site during 284
SOAS, P[NO3] was used as a proxy for NO3 radicals present in the atmosphere to examine if there 285
is any association of it with isoprene-derived SOA tracers. 286
3. Results and Discussion 287
3.1. Overview of the study 288
The campaign extended from June 1 through July 16, 2013. Temperature during this period 289
ranged from a high of 32.6 °C to a low of 20.5 °C, with an average of ~26.4 °C. RH varied from 290
37-96% throughout the campaign, with an average of 71.5%. Rainfall occurred intermittently over 291
2-3 day periods and averaged 0.1 inches per day. Wind analysis reveals that air masses approached 292
largely from the south-southeast at an average wind speed of 2 m s-1. Summaries of meteorological 293
conditions as well as wind speed and direction during the course of the campaign are given in 294
Table 2 and illustrated in Figures 1 and 2. 295
The average concentration of carbon monoxide (CO), a combustion byproduct, was 208.7 296
ppbv. The mean concentration of O3 was significantly higher (t-test, p-value < 0.05) on intensive 297
sampling days (37.0 ppbv) than regular sampling days (25.2 ppbv). Concentrations of NOx, NH3, 298
and SO2 were lower averaging 7.8, 1.9, and 0.9 ppbv, respectively. On average, OC and WSOC 299
levels were 7.2 (n = 120) and 4 µg m-3 (n = 100), respectively. The largest inorganic component of 300
PM2.5 was SO42-, which averaged 2 μg m-3 with excursions between 0.4 and 4.9 μg m-3 during the 301
campaign. NH4+ and NO3- were present at low levels, averaging 0.66 and 0.14 μg m-3, respectively. 302
Time series of gas and PM2.5 components are shown in Figure 2. WSOC accounted for 35% of OC 303
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
15
mass (Figure S3a), and was smaller than that recently reported in rural areas during SOAS 304
(Budisulistiorini et al., 2015; Hu et al., 2015), but consistent with previous observations at the 305
BHM site (Ding et al., 2008). WSOC/OC ratios are commonly lower in urban than rural areas, as 306
a consequence of higher primary OC emissions; thus, PM at BHM probably contains increased 307
OC. 308
Diurnal variation of meteorological parameters, trace gases, and PM2.5 components are 309
shown in Figure S4 of the Supplement. Temperature dropped during nighttime, and reached a 310
maximum in the afternoon (Figure S4a). Conversely, RH was low during day and high at night. 311
High-NOx levels were found in the early morning and decreased during the course of the day 312
(Figure S4c), most likely in conjunction with rising O3 levels. O3 reached a maximum 313
concentration between 12 - 3 pm due to photochemistry (Figure S4b). SO2 was slightly higher in 314
the morning (Figure S4c), but decreased during the day most likely as a result of planetary 315
boundary layer (PBL) dynamics. NH3 remained fairly constant throughout the day (Figure S4c). 316
No significant diurnal variation was found in the concentration of inorganic PM2.5 components, 317
including SO42-, NO3-, and NH4+ (Figure S4d). Unfortunately, a measurement of isoprene could 318
not be made at BHM during the campaign. However, the diurnal trend of isoprene levels might be 319
similar to the data at the CTR site (Xu et al., 2015), which is only 61 miles away from BHM. Xu 320
et al. (2015) observed the highest levels of isoprene (~ 6 ppb) at CTR in the mid-afternoon (3 pm 321
local time) and its diurnal trend was similar to isoprene-OA measured by the Aerodyne Aerosol 322
Mass Spectrometer (AMS) during the SOAS campaign. 323
3.2 Characterization of Isoprene SOA 324
Table 3 summarizes the mean and maximum concentrations of known isoprene-derived 325
SOA tracers detected by GC/EI-MS and UPLC/ESI-HR-QTOFMS. Levoglucosan was also 326
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
16
analyzed as a tracer for biomass burning. Among the isoprene-derived SOA tracers, the highest 327
mean concentration was for 2-methyltetrols (376 ng m-3), followed by the sum of C5-alkene triols 328
(181 ng m-3) and the IEPOX-derived OS (165 ng m-3). The concentrations account for 3.8%, 1.8% 329
and 1.6%, respectively, of total OM mass. Noteworthy is that maximum concentrations of 2-330
methylerythritol (a 2-methyltetrol isomer; 1049 ng m-3), IEPOX-derived OS (865 ng m-3) and (E)-331
2-methylbut-3-ene-1,2,4-triol (879 ng m-3) were attained during the intensive sampling period 4-7 332
pm local time on June 15, 2013, following five consecutive days of dry weather (Figure 2a and 333
2d) when high levels of isoprene, SO42-, and NOx were forecast. 334
Together, the IEPOX-derived SOA tracers, which represent SOA formation from isoprene 335
oxidation predominantly under the low-NOx pathway, comprised 92.5% of the total detected 336
isoprene-derived SOA tracer mass at the BHM site. This contribution is slightly lower than 337
observations reported at rural sites located in Yorkville, GA (97.5%) and Look Rock, Tennessee 338
(LRK) (97%) (Lin et al., 2013b; Budisulistiorini et al., 2015). 339
The sum of MAE/HMML-OS and 2-MG, which represent SOA formation from isoprene 340
oxidation predominantly under the high-NOx pathway, contributed 3.25% of the total isoprene-341
derived SOA tracer mass, while the OS derivative of glycolic acid (GA sulfate) contributed 3.3%. 342
The contribution of GA sulfate was consistent with the level of GA sulfate measured by the 343
airborne NOAA Particle Analysis Laser Mass Spectrometer (PALMS) over the continental U.S. 344
during the Deep Convective Clouds and Chemistry Experiment and SEAC4RS (Liao et al., 2015). 345
However, the contribution of GA sulfate to the total OM at BHM (0.3%) is lower than aircraft-346
based measurements made by Liao et al. (2015) near the ground in the eastern U.S. (0.9%). GA 347
sulfate can form from biogenic and anthropogenic emissions other than isoprene, including 348
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
17
glyoxal, which is thought to be a primary source of GA sulfate (Galloway et al., 2009). For this 349
reason, GA sulfate will not be further discussed in this study. 350
Isoprene SOA contribution to total OM was estimated by assuming the OM/OC ratio 1.6 351
based on the recent studies (El-Zanan et al., 2009; Simon et al., 2011; Ruthenburg et al., 2014; 352
Blanchard et al., 2015). On average, isoprene-derived SOA tracers (sum of both IEPOX- and 353
MAE/HMML-derived SOA tracers) contributed ~7% (ranging to ~ 20% at times) of the total 354
particulate OM mass. The average contribution is lower than measured at other sites in the S.E. 355
USA, including both rural LRK, (Budisulistiorini et al., 2015; Hu et al., 2015) and urban Atlanta, 356
GA (Budisulistiorini et al., 2013). The contribution of SOA tracers to OM in the current study was 357
estimated on the basis of offline analysis of filters, while tracer estimates in the two earlier studies 358
was based on online ACSM/AMS measurements. The low isoprene SOA/OM ratio is consistent 359
with the low WSOC/OC reported in section 3.1, suggesting an increased contribution of primary 360
OA or secondary OM to the total OM at BHM. However, it should be noted that total IEPOX-361
derived SOA mass at BHM may actually be closer to ~14% since recent measurements by the 362
Aerodyne ACSM at LRK indicated that tracers could only account for ~50% of the total IEPOX-363
derived SOA mass resolved by the ACSM (Budisulistiorini et al., 2015). Unfortunately, an 364
Aerodyne ACSM or AMS was not available at the BHM site, precluding confirmation of that 365
IEPOX-derived SOA mass at BHM might account for 14% (on average) of the total OM mass. 366
Levoglucosan, a biomass-burning tracer, averaged 1% of total OM with spikes up to 8%, the same 367
level measured for 2-methylthreitol and (E)-2-methylbut-3-ene-1,2,4-triol (Table 3). The ratio of 368
average levoglucosan at BHM relative to CTR was 5.4 suggesting significantly more biomass 369
burning impacting the BHM site. 370
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
18
IEPOX- and MAE/HMML-derived SOA tracers accounted for 18% and 0.4% of the 371
WSOC mass, respectively (Figure S3b), lower than the respective contributions of 24% and 0.7% 372
measured at LRK (Budisulistiorini et al., 2015). 373
Figure S5 shows no diurnal variation for the average day and night concentrations of 374
isoprene-derived SOA tracers. Thus cooler nighttime temperatures also do not appear to enhance 375
gas-to-particle partitioning at the BHM site. Figures S6 and S7 show the variation of isoprene-376
derived SOA tracers during intensive sampling periods. The highest concentrations were usually 377
observed in samples collected from 4 pm – 7 pm, local time; however, no statistical significance 378
were observed between intensive periods. This observation illustrates the importance of the higher 379
time-resolution of the tracer data during intensive sampling periods over course of the campaign 380
(Table S2-S6). An additional consequence of the intensive sampling periods was resolution of a 381
significant correlation between isoprene SOA tracers and O3 to be discussed in more detail in 382
section 3.3.2. 383
3.3 Influence of anthropogenic emissions on isoprene-derived SOA 384
3.3.1 Effects of reactive nitrogen-containing species 385
During the campaign, no isoprene-derived SOA tracers, including MAE/HMML-derived 386
OS and 2-MG, correlated with NOx or NOy (r2 = 0, n = 120). This is inconsistent with the current 387
understanding of SOA formation from isoprene oxidation pathways under high-NOx conditions, 388
which proceeds through uptake of MAE (Lin et al., 2013a), and, as recently suggested, HMML 389
(Nguyen et al., 2015), to yield 2-MG and its OS derivative. Plume age, as a ratio of NOx:NOy,, in 390
this study was highly correlated with O3 (r2 = 0.79, n = 120). This correlation might be explained 391
by the photolysis of NO2, which is abundant due to traffic at the urban ground site, resulting in 392
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
19
formation of tropospheric O3. A negative correlation coefficient (r = - 0.47, n = 120) between 393
plume age and 2-MG abundance was found, suggesting that formation of some 2-MG may be 394
associated with ageing of air masses. 395
A previous study supported a major role for NO3 in the nighttime chemistry of isoprene 396
(Ng et al., 2008). Correlation of IEPOX- and MAE/HMML-derived SOA with nighttime NO2, O3, 397
and P[NO3] were examined in this study (Figures 3 and 4). As shown in Figure 3f, a moderate 398
correlation between MAE/HMML-derived SOA and nighttime P[NO3] (r2 = 0.57, n = 40) was 399
observed. The regression analysis revealed a significant correlation at the 95% confidence interval 400
(p-value < 0.05) (Table S7). This finding suggests that some MAE/HMML-derived SOA may form 401
locally from the reaction of isoprene with NO3 radical at night. A field study reported a peak 402
isoprene mixing ratio in early evening (Starn et al., 1998) as the PBL height decreases at night. As 403
a result, lowering PBL heights could concentrate the remaining isoprene, NO2, and O3 that can 404
continue to react during the course of the evening. 2-MG formation has been reported to be NO2-405
dependent via the formation and further oxidation of MPAN (Surratt et al., 2006; Chan et al., 406
2010). Hence, decreasing PBL may be related to nighttime MAE/HMML-derived SOA formation 407
through isoprene oxidation by both P[NO3] and NO2. 408
Although P[NO3] depends on both NO2 and O3 levels, O3 correlates moderately with 409
MAE/HMML-derived SOA tracers during day (r2 = 0.48, n = 75), but not at night (r2 = 0.08, n = 410
45). The effect of O3 on isoprene-derived SOA formation during daytime will be discussed further 411
in section 3.3.2. NO2 levels correlate only weakly with MAE/HMML-derived SOA tracers, (r2 = 412
0.26, n = 45) indicating that NO2 levels alone do not explain the moderate correlation of P[NO3] 413
with these tracers. To our knowledge, correlation of P[NO3] with high-NOx SOA tracers has not 414
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
20
been observed in previous field studies., indicating that further work is needed to examine the 415
potential role of nighttime NO3 radicals in forming these SOA tracers. 416
As shown in Figure 4f, IEPOX-derived SOA was weakly correlated (r2 = 0.26, n = 40) with 417
nighttime P[NO3]. The correlation appears to be driven by the data at the low end of the scale and 418
could therefore be misleading. However, Schwantes et al. (2015) demonstrated that NO3-initaited 419
oxidation of isoprene yields isoprene nitrooxy hydroperoxides (INEs) through nighttime reaction: 420
RO2 + HO2, which on further oxidation yielded isoprene nitrooxy hydroxyepoxides (INHEs). The 421
INHEs undergo reactive uptake onto acidic sulfate aerosol to yield SOA constituents similar to 422
those of IEPOX-derived SOA. The present study raises the possibility that a fraction of IEPOX-423
derived SOA comes from NO3-initiated oxidation of isoprene at night. The work of Ng et al. 424
(2008) does not explain the weak association we observe here between IEPOX-derived SOA 425
tracers and P[NO3] as a consequence of the reactions RO2 + RO2 and RO2 + NO3 reactions 426
dominating in those experiments. It is now thought that RO2 + HO2 should dominate in field studies 427
(Schwantes et al., 2015; Paulot et al., 2009). 428
3.3.2 Effect of O3
429
During the daytime, O3 was moderately correlated (r2 = 0.48, n = 75) with total 430
MAE/HMML-derived SOA (Figure 3b). This correlation was stronger (r2 = 0.72, n = 30, p-value 431
< 0.05, Table S7) when filters taken during regular daytime sampling periods are considered, 432
suggesting that formation of MACR (a precursor to MAE and HMML) (Lin et al., 2013b; Nguyen 433
et al., 2015) was enhanced by oxidation of isoprene by O3 (Kamens et al., 1982). O3 was not 434
correlated (r2 = 0.08, n = 45) with MAE/HMML-derived SOA at night (Figure 3e). The latter 435
finding is consistent with the absence of photolysis to drive the production of O3. However, 436
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
21
residual O3 may play an important role at night to form MAE/HMML-derived SOA via the P[NO3] 437
pathway discussed in section 3.3.1. 438
O3 was not correlated (r2= 0.10, n = 75) with IEPOX-derived SOA during daytime (Figure 439
4b), but weakly correlated with 2-methylerythritol (r2 = 0.25, n = 30) as shown in Table S2, 440
especially during intensive 3 sampling periods (r2 = 0.34, n = 15, Table S5). An important 441
observation with regard to this result is that no correlation has been found between O3 and 2-442
methyltetrols (r2 < 0.01) in previous field studies (Lin et al., 2013b; Budisulistiorini et al., 2015). 443
Isoprene ozonolysis yielded 2-methyltetrols in chamber studies in the presence of acidified sulfate 444
aerosol (Riva et al., 2015) but C5-alkene-triols were not formed by this pathway. The greatest 445
abundance of isoprene-derived SOA tracers in daytime samples was generally observed in 446
intensive 3 samples; however, there was no statistical significance observed between intensive 447
samples. The moderate correlation (r2 = 0.34, n = 15, p-value < 0.05) between O3 and the 2-448
methyltetrols observed in intensive 3 samples occurred when O3 reached maximum levels, 449
suggesting that ozonolysis of isoprene plays a role in 2-methyltetrol formation. Lack of correlation 450
between O3 and C5-alkene triols during intensive 3 sampling (r2 = 0.10, n = 15) supports this 451
contention. A putative pathway is formation of hydroperoxides that partition to wet acidic sulfate 452
aerosols and react further to yield 2-methyltetrols. Additional work using authentic standards is 453
needed to validate this tentative hypothesis. 454
3.3.3 Effect of particle SO42-
455
SO42- was moderately correlated with IEPOX-derived SOA (r2 = 0.36, n = 117) and 456
MAE/HMML-derived SOA (r2 = 0.33, n = 117) at the 95% confidence interval as shown in Table 457
S7. The strength of the correlations was consistent with studies at other sites across the 458
Southeastern U.S. (Budisulistiorini et al., 2013; Lin et al., 2013b; Budisulistiorini et al., 2015; Xu 459
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
22
et al., 2015). Aerosol surface area provided by acidic SO42- has been demonstrated to control the 460
uptake of isoprene-derived epoxides (Lin et al., 2012; Gaston et al., 2014; Nguyen et al., 2014; 461
Riedel et al., 2015). 462
Furthermore, SO42- is proposed to enhance IEPOX-derived SOA formation by providing 463
particle water (H2Optcl) required for IEPOX uptake (Xu et al., 2015). Aerosol SO42- also promotes 464
acid-catalyzed ring-opening reactions of IEPOX by H+, proton donors such as NH4+, and 465
nucleophiles (e.g., H2O, SO42-, or NO3-) (Surratt et al., 2010; Nguyen et al., 2014). Since SO42- 466
tends to drive both particle water and acidity (Fountoukis and Nenes, 2007), the extent to which 467
each influences isoprene SOA formation during field studies remains unclear. Multivariate linear 468
regression analysis on SOAS data from the CTR site and the SCAPE dataset revealed a statistically 469
significant positive linear relationship between SO42- and the isoprene (IEPOX)-OA factor 470
resolved by positive matrix factorization (PMF). On the basis of this analysis the abundance of 471
SO42- was concluded to control directly the isoprene SOA formation over broad areas of the 472
Southeastern U.S. (Xu et al., 2015), consistent with previous reports (Lin et al., 2013; 473
Budisulistiorini et al., 2013; Budisulistiorini et al., 2015). Another potential pathway for SO42-
474
levels to enhance isoprene SOA formation is through salting-in effects; however, systematic 475
investigations of this effect are lacking and further studies are warranted (Xu et al., 2015). 476
3.3.4 Effect of aerosol acidity 477
The aerosol at BHM was acidic throughout the SOAS campaign (pH range 1.60 1.94, 478
average 1.76) in accord with a study by Guo et. al. (2014) that found aerosol pH ranging from 479
0 2 throughout the southeastern U.S. However, no correlation of pH with isoprene SOA 480
formation was observed at BHM, also consistent with previous findings using the thermodynamic 481
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
23
models to estimate aerosol acidity in many field sites across the southeastern U.S. region, including 482
Yorkville, GA (YRK) (Lin et al., 2013b), Jefferson Street, GA (JST) (Budisulistiorini et al., 2013), 483
and LRK (Budisulistiorini et al., 2015). However, it is important to point out that the lack of 484
correlation between SOA tracers and acidity may stem from the small variations in aerosol acidity 485
throughout the campaign. Gaston et al. (2014) and Riedel et al. (2015) recently demonstrated that 486
an aerosol pH < 2 at atmospherically-relevant aerosol surface areas would allow reactive uptake 487
of IEPOX onto acidic (wet) sulfate aerosol surfaces to be competitive with other loss processes 488
(e.g., deposition and reaction of IEPOX with OH). In fact, it was estimated that under such 489
conditions IEPOX would have a lifetime of ~ 5 hr. The constant presence of acidic aerosol has 490
also been observed at other field sites in the southeastern U.S. (Budisulistiorini et al., 2013; 491
Budisulistiorini et al., 2015; Xu et al., 2015), supporting a conclusion that acidity is not the limiting 492
variable in forming isoprene SOA. 493
3.4 Comparison among different sampling sites during 2013 SOAS campaign 494
Table 5 summarizes the mean concentration and contribution of each isoprene SOA tracer 495
at BHM, CTR, and LRK. BHM is an industrial-residential area, LRK and CTR are rural areas, 496
although LRK is influenced by a diurnal upslope/downslope cycle of air from an urban locality 497
(Knoxville) (Tanner et al., 2005). IEPOX-derived SOA was predominant at all three sites during 498
the SOAS campaign, while MAE/HMML-derived SOA constituted a minor contribution. The 499
average ratio of 2-methyltetrols to C5-alkene triols at BHM was 2.2, nearly double that of CTR 500
(1.3) and LRK (1.1). Although 2-methyltetrols and C5-alkene triols are considered to form readily 501
from the acid-catalyzed reactive uptake and multiphase chemistry of IEPOX (Edney et al., 2005; 502
Surratt et al., 2006), Riva et al. (2015) recently demonstrated that only 2-methyltetrols can be 503
formed via isoprene ozonolysis in the presence of acidic sulfate aerosol. The higher levels of the 504
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
24
2-methyltetrols observed at the urban BHM site indicates a likely competition between the IEPOX 505
uptake and ozonolysis pathways. Together, these findings suggest that urban O3 may play an 506
important role in forming the 2-methyltetrols observed at BHM. There were notable trends found 507
among the three sites: (1) average C5-alkene triol concentrations were higher at CTR (214.1 ng m-
508
3) than at BHM (169.7 ng m-3) and LRK (144.4 ng m-3); (2) average isomeric 3-MeTHF-diol 509
concentrations were lower at CTR (0.2 ng m-3) than the BHM (15.4 ng m-3) or LRK (4.4 ng m-3) 510
sites. Except for the 2-methyltetrols, reasons for the differences observed for the other tracers 511
between sites remains unclear and warrant future investigations. 512
513
4. Conclusions 514
This study examined isoprene SOA tracers in PM2.5 samples collected at the BHM ground 515
site during the 2013 SOAS campaign and revealed the complexity and potential multitude of 516
chemical pathways leading to isoprene SOA formation. Isoprene SOA contributed up to ~20% 517
(~7% on average) of total OM mass. IEPOX-derived SOA tracers were responsible for 92% of the 518
total quantified isoprene SOA tracer mass, with 2-methyltetrols being the major component (47%). 519
Differences in the relative contributions of IEPOX- and MAE/HMML-derived SOA tracers at 520
BHM and the rural CTR and LRK sites (Budisulistiorini et al., 2015) during the 2013 SOAS 521
campaign, support suggestions that anthropogenic emissions effect isoprene SOA formation. The 522
correlation between 2-methyltetrols and O3 at BHM is in accord with work by Riva et al. (2015), 523
demonstrating a potential role of O3 in generating isoprene-derived SOA in addition to the 524
currently accepted IEPOX multiphase pathway. 525
At BHM, the statistical correlation of particulate SO42- with IEPOX- (r2 = 0.36, n = 117, p 526
< 0.05) and MAE-derived SOA tracers (r2 = 0.33, n = 117, p < 0.05) suggests that SO42- plays a 527
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
25
role in isoprene SOA formation. Although none of isoprene-derived SOA tracers correlated with 528
gas-phase NOx and NOy, MAE/HMML-derived SOA tracers correlated with nighttime P[NO3] (r2 529
= 0.57, n = 400), indicating that NO3 may affect local MAE/HMML-derived SOA formation. 530
Nighttime P[NO3] was weakly correlated (r2 = 0.26, n = 40) with IEPOX-derived SOA tracers, 531
lending some support to recent work by Schwantes et al. (2015) showing that isoprene + NO3 532
yields INHEs that can by undergo reactive uptake to yield IEPOX tracers and contribute to IEPOX-533
derived SOA tracer loadings. In addition, nighttime 2-methyltetrol levels in the urban atmosphere 534
deviate from the conventional understanding of isoprene SOA formation in terms of segregated 535
NOx dependent regimes. The correlation of daytime O3 with MAE/HMML-derived SOA and with 536
2-methyltetrols offers a new insight into influences on isoprene SOA formation. Notably, O3 has 537
not been reported to correlate with isoprene-derived SOA tracers in previous field studies (Lin et 538
al., 2013b; Budisulistiorini et al., 2015). In this study, the strong correlation (r2 = 0.72, n = 30) at 539
the 95% confidence interval of O3 with MAE/HMML-derived SOA tracers during the regular 540
daytime sampling schedule indicates that O3 likely oxidizes some isoprene to MACR as precursor 541
of 2-MG at BHM. The weak correlation (r2 = 0.16, n = 75) between O3 and 2-methyltetrols early 542
in the day as well as the better correlation (r2 = 0.34, n = 15) later in the day (intensive 3, 4-7 PM 543
local time) are consistent with recent laboratory studies demonstrating that 2-methyltetrols can be 544
formed via isoprene ozonolysis in the presence of acidified sulfate aerosol (Riva et al., 2015). 545
Although urban O3 and nighttime P[NO3] may have a role in local formation of 546
MAE/HMML- and IEPOX-derived SOA tracers at BHM, this does not appear to explain the 547
majority of the SOA tracers, since no significant day-night variation of the entire group of tracers 548
was observed during the campaign. The majority of IEPOX-derived SOA was likely formed when 549
isoprene SOA precursors (IEPOX) were generated upwind and transported to the BHM site. Wind 550
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
26
directions during the campaign are consistent with long-range transport of isoprene SOA 551
precursors from southwest of the site, which is covered by forested areas. The absence of a 552
correlation of aerosol acidity with MAE/HMML- and IEPOX-derived SOA tracers indicates that 553
acidity is not the limiting variable that controls formation of these compounds. However, the lack 554
of correlation between SOA tracers and acidity may stem from nearly invariant aerosol acidity 555
throughout the campaign. Hence, despite laboratory studies demonstrating that aerosol acidity can 556
enhance isoprene SOA formation (Surratt et al., 2007; Surratt et al., 2010; Lin et al., 2012), the 557
effect may not be significant in the southeastern U.S. during the summer months due to the constant 558
acidity of aerosols. Future work should examine how well current models can predict the isoprene 559
SOA levels observed during this study, especially since urban emissions are directly present. 560
Furthermore, explicit models are now available to predict the isoprene SOA tracers measured here 561
(McNeill et al., 2012; Pye et al., 2013), which will allow the modeling community to test the 562
current parameterizations that are used to capture the enhancing effect of anthropogenic pollutants 563
on isoprene-derived SOA formation. In addition, the significant correlations of isoprene-derived 564
SOA tracers with P[NO3] observed during this study indicate a need to better understand nighttime 565
chemistry of isoprene. Lastly, although O3 appears to have an enhancing effect on isoprene-566
derived SOA tracers, the intermediates are unknown. Hydroperoxides suggested by Riva et al. 567
(2015) may be key, but chamber experiments with authentic precursors are needed to test this 568
hypothesis. 569
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
27
Acknowledgements 570
This work was funded by the U.S. Environmental Protection Agency (EPA) through grant number 571
835404. The contents of this publication are solely the responsibility of the authors and do not 572
necessarily represent the official views of the U.S. EPA. Further, the U.S. EPA does not endorse 573
the purchase of any commercial products or services mentioned in the publication. The authors 574
would also like to thank the Electric Power Research Institute (EPRI) for their support. This study 575
was supported in part by the National Oceanic and Atmospheric Administration (NOAA) Climate 576
Program Office’s AC4 program, award number NA13OAR4310064. The authors thank the 577
Camille and Henry Dreyfus Postdoctoral Fellowship Program in Environmental Chemistry for 578
their financial support. The authors thank Louisa Emmons and Christoph Knote for their assistance 579
with chemical forecasts made available during the SOAS campaign. We would like to thank 580
Annmarie Carlton, Joost deGouw, Jose Jimenez, and Allen Goldstein for helping to organize the 581
SOAS campaign and coordinating communication between ground sites. UPLC/ESI-HR-Q-582
TOFMS analyses were conducted in the UNC-CH Biomarker Mass Facility located within the 583
Department of Environmental Sciences and Engineering, which is a part of the UNC-CH Center 584
for Environmental Health and Susceptibility supported by National Institute for Environmental 585
Health Sciences (NIEHS), grant number 5P20-ES10126. WSOC measurements at the University 586
of Iowa were supported through EPA STAR grant 8354101. The authors thank Theran Riedel for 587
useful discussions. We also thank SCG Chemicals Co., Ltd., Siam Cement Group, Thailand, for 588
the full support for W. Rattanavaraha attending UNC, Chapel Hill. 589
590
591
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
28
References 592
Birch, M. E., and Cary, R. A.: Elemental carbon-based method for occupational monitoring of 593
particulate diesel exhaust: methodology and exposure issues, Analyst, 121, 1183-1190, 594
1996. 595
Blanchard, C. L., Hidy, G. M., Shaw, S., Baumann, K., and Edgerton, E. S.: Effects of emission 596
reductions on organic aerosol in the southeastern United States, Atmos. Chem. Phys. 597
Discuss., 15, 17051-17092, doi:10.5194/acpd-15-17051-2015, 2015. 598
Boucher, O., Randall, D., Artaxo, P., Bretherton, C., Feingold, G., Forster, P., Kerminen, V.-M., 599
Kondo, Y., Liao, H., and Lohmann, U.: Clouds and aerosols, in: Climate change 2013: 600
the physical science basis. Contribution of Working Group I to the Fifth Assessment 601
Report of the Intergovernmental Panel on Climate Change, Cambridge University Press, 602
571-657, 2013. 603
Budisulistiorini, S., Li, X., Bairai, S., Renfro, J., Liu, Y., Liu, Y., McKinney, K., Martin, S., 604
McNeill, V., and Pye, H.: Examining the effects of anthropogenic emissions on isoprene-605
derived secondary organic aerosol formation during the 2013 Southern Oxidant and 606
Aerosol Study (SOAS) at the Look Rock, Tennessee, ground site, Atmospheric Chemistry 607
and Physics Discussions, 15, 7365-7417, 2015. 608
Budisulistiorini, S. H., Canagaratna, M. R., Croteau, P. L., Marth, W. J., Baumann, K., Edgerton, 609
E. S., Shaw, S. L., Knipping, E. M., Worsnop, D. R., and Jayne, J. T.: Real-time 610
continuous characterization of secondary organic aerosol derived from isoprene 611
epoxydiols in downtown Atlanta, Georgia, using the Aerodyne Aerosol Chemical 612
Speciation Monitor, Environ. Sci. Technol., 47, 5686-5694, 2013. 613
Carlton, A., Wiedinmyer, C., and Kroll, J.: A review of Secondary Organic Aerosol (SOA) 614
formation from isoprene, Atmospheric Chemistry and Physics, 9, 4987-5005, 2009. 615
Carlton, A. G., Bhave, P. V., Napelenok, S. L., Edney, E. O., Sarwar, G., Pinder, R. W., Pouliot, 616
G. A., and Houyoux, M.: Model representation of secondary organic aerosol in CMAQv4. 617
7, Environ. Sci. Technol., 44, 8553-8560, 2010a. 618
Carlton, A. G., Pinder, R. W., Bhave, P. V., and Pouliot, G. A.: To what extent can biogenic SOA 619
be controlled?, Environ. Sci. Technol., 44, 3376-3380, 2010b. 620
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
29
Chan, A., Chan, M., Surratt, J., Chhabra, P., Loza, C., Crounse, J., Yee, L., Flagan, R., Wennberg, 621
P., and Seinfeld, J.: Role of aldehyde chemistry and NOx concentrations in secondary 622
organic aerosol formation, Atmospheric Chemistry and Physics, 10, 7169-7188, 2010. 623
Claeys, M., Graham, B., Vas, G., Wang, W., Vermeylen, R., Pashynska, V., Cafmeyer, J., Guyon, 624
P., Andreae, M. O., and Artaxo, P.: Formation of secondary organic aerosols through 625
photooxidation of isoprene, Science, 303, 1173-1176, 2004. 626
Ding, X., Zheng, M., Yu, L., Zhang, X., Weber, R. J., Yan, B., Russell, A. G., Edgerton, E. S., and 627
Wang, X.: Spatial and seasonal trends in biogenic secondary organic aerosol tracers and 628
water-soluble organic carbon in the southeastern United States, Environ. Sci. Technol., 629
42, 5171-5176, 2008. 630
Edgerton, E. S., Hartsell, B. E., Saylor, R. D., Jansen, J. J., Hansen, D. A., and Hidy, G. M.: The 631
Southeastern Aerosol Research and Characterization Study, part 3: Continuous 632
measurements of fine particulate matter mass and composition, Journal of the Air & 633
Waste Management Association, 56, 1325-1341, 2006. 634
Edney, E. O., Kleindienst, T. E., Jaoui, M., Lewandowski, M., Offenberg, J. H., Wang, W., and 635
Claeys, M.: Formation of 2-methyl tetrols and 2-methylglyceric acid in secondary organic 636
aerosol from laboratory irradiated isoprene/NOX/SO2/air mixtures and their detection in 637
ambient PM2.5 samples collected in the eastern United States, Atmospheric Environment, 638
39, 5281-5289, http://dx.doi.org/10.1016/j.atmosenv.2005.05.031, 2005. 639
El-Zanan, H. S., Zielinska, B., Mazzoleni, L. R., and Hansen, D. A.: Analytical determination of 640
the aerosol organic mass-to-organic carbon ratio, Journal of the Air & Waste Management 641
Association, 59, 58-69, 2009. 642
Emmons, L. K., Walters, S., Hess, P. G., Lamarque, J. F., Pfister, G. G., Fillmore, D., Granier, C., 643
Guenther, A., Kinnison, D., Laepple, T., Orlando, J., Tie, X., Tyndall, G., Wiedinmyer, 644
C., Baughcum, S. L., and Kloster, S.: Description and evaluation of the Model for Ozone 645
and Related chemical Tracers, version 4 (MOZART-4), Geosci. Model Dev., 3, 43-67, 646
10.5194/gmd-3-43-2010, 2010. 647
Foley, K., Roselle, S., Appel, K., Bhave, P., Pleim, J., Otte, T., Mathur, R., Sarwar, G., Young, J., 648
and Gilliam, R.: Incremental testing of the Community Multiscale Air Quality (CMAQ) 649
modeling system version 4.7, Geoscientific Model Development, 3, 205-226, 2010. 650
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
30
Fountoukis, C., and Nenes, A.: ISORROPIA II: a computationally efficient thermodynamic 651
equilibrium model for K+–Ca 2+–Mg 2+–NH 4+–Na+–SO 4 2−–NO 3−–Cl−–H 2 O 652
aerosols, Atmospheric Chemistry and Physics, 7, 4639-4659, 2007. 653
Fountoukis, C., Nenes, A., Sullivan, A., Weber, R., Reken, T. V., Fischer, M., Matias, E., Moya, 654
M., Farmer, D., and Cohen, R.: Thermodynamic characterization of Mexico City aerosol 655
during MILAGRO 2006, Atmospheric Chemistry and Physics, 9, 2141-2156, 2009. 656
Galloway, M. M., Chhabra, P. S., Chan, A. W. H., Surratt, J. D., Flagan, R. C., Seinfeld, J. H., and 657
Keutsch, F. N.: Glyoxal uptake on ammonium sulphate seed aerosol: reaction products 658
and reversibility of uptake under dark and irradiated conditions, Atmos. Chem. Phys., 9, 659
3331-3345, 10.5194/acp-9-3331-2009, 2009. 660
Graham, R. A., and Johnston, H. S.: The photochemistry of the nitrate radical and the kinetics of 661
the nitrogen pentoxide-ozone system, The Journal of Physical Chemistry, 82, 254-268, 662
1978. 663
Guenther, A. B., Jiang, X., Heald, C. L., Sakulyanontvittaya, T., Duhl, T., Emmons, L. K., and 664
Wang, X.: The Model of Emissions of Gases and Aerosols from Nature version 2.1 665
(MEGAN2.1): an extended and updated framework for modeling biogenic emissions, 666
Geosci Model Dev, 5, 1471-1492, 10.5194/gmd-5-1471-2012, 2012. 667
Grieshop, A., P., Logue, J., M., Donahue., J., M., and Robinson, A., L.: Laboratory investigation 668
of photochemical oxidation of organic aerosol from wood fires 1 : measurement and 669
simulation of organic aerosol evolution, Atmos. Chem. Phys., 9, 1263-1277, 2009. 670
Hallquist, M., Wenger, J. C., Baltensperger, U., Rudich, Y., Simpson, D., Claeys, M., Dommen, 671
J., Donahue, N. M., George, C., Goldstein, A. H., Hamilton, J. F., Herrmann, H., 672
Hoffmann, T., Iinuma, Y., Jang, M., Jenkin, M. E., Jimenez, J. L., Kiendler-Scharr, A., 673
Maenhaut, W., McFiggans, G., Mentel, T. F., Monod, A., Prévôt, A. S. H., Seinfeld, J. 674
H., Surratt, J. D., Szmigielski, R., and Wildt, J.: The formation, properties and impact of 675
secondary organic aerosol: current and emerging issues, Atmos. Chem. Phys., 9, 5155-676
5236, 10.5194/acp-9-5155-2009, 2009. 677
Hansen, D. A., Edgerton, E. S., Hartsell, B. E., Jansen, J. J., Kandasamy, N., Hidy, G. M., and 678
Blanchard, C. L.: The Southeastern aerosol research and characterization study: part 1—679
overview, Journal of the Air & Waste Management Association, 53, 1460-1471, 2003. 680
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
31
Henze, D. K., Seinfeld, J. H., and Shindell, D. T.: Inverse modeling and mapping US air quality 681
influences of inorganic PM 2.5 precursor emissions using the adjoint of GEOS-Chem, 682
Atmospheric Chemistry and Physics, 9, 5877-5903, 2009. 683
Herron, J. T., and Huie, R. E.: Rate constants for the reactions of ozone with ethene and propene, 684
from 235.0 to 362.0. deg. K, The Journal of Physical Chemistry, 78, 2085-2088, 1974. 685
Hu, W., Campuzano-Jost, P., Palm, B., Day, D., Ortega, A., Hayes, P., Krechmer, J., Chen, Q., 686
Kuwata, M., and Liu, Y.: Characterization of a real-time tracer for Isoprene Epoxydiols-687
derived Secondary Organic Aerosol (IEPOX-SOA) from aerosol mass spectrome-ter 688
measurements, Atmospheric Chemistry and Physics Discussions, 15, 11223-11276, 2015. 689
Kamens, R., Gery, M., Jeffries, H., Jackson, M., and Cole, E.: Ozone-isoprene reactions: product 690
formation and aerosol potential, Int. J. Chem. Kinet., 14, 955-975, 1982. 691
Kanakidou, M., Seinfeld, J. H., Pandis, S. N., Barnes, I., Dentener, F. J., Facchini, M. C., Van 692
Dingenen, R., Ervens, B., Nenes, A., Nielsen, C. J., Swietlicki, E., Putaud, J. P., 693
Balkanski, Y., Fuzzi, S., Horth, J., Moortgat, G. K., Winterhalter, R., Myhre, C. E. L., 694
Tsigaridis, K., Vignati, E., Stephanou, E. G., and Wilson, J.: Organic aerosol and global 695
climate modelling: a review, Atmos. Chem. Phys., 5, 1053-1123, 10.5194/acp-5-1053-696
2005, 2005. 697
Karambelas, A., Pye, H. O., Budisulistiorini, S. H., Surratt, J. D., and Pinder, R. W.: Contribution 698
of isoprene epoxydiol to urban organic aerosol: evidence from modeling and 699
measurements, Environmental Science & Technology Letters, 1, 278-283, 2014. 700
Kroll, J. H., Ng, N. L., Murphy, S. M., Flagan, R. C., and Seinfeld, J. H.: Secondary organic aerosol 701
formation from isoprene photooxidation under high-NOx conditions, Geophysical 702
Research Letters, 32, 2005. 703
Kroll, J. H., Ng, N. L., Murphy, S. M., Flagan, R. C., and Seinfeld, J. H.: Secondary Organic 704
Aerosol Formation from Isoprene Photooxidation, Environ. Sci. Technol., 40, 1869-1877, 705
10.1021/es0524301, 2006. 706
Liao, J., Froyd, K. D., Murphy, D. M., Keutsch, F. N., Yu, G., Wennberg, P. O., St Clair, J. M., 707
Crounse, J. D., Wisthaler, A., and Mikoviny, T.: Airborne measurements of 708
organosulfates over the continental US, Journal of Geophysical Research: Atmospheres, 709
120, 2990-3005, 2015. 710
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
32
Lin, Y.-H., Zhang, Z., Docherty, K. S., Zhang, H., Budisulistiorini, S. H., Rubitschun, C. L., Shaw, 711
S. L., Knipping, E. M., Edgerton, E. S., and Kleindienst, T. E.: Isoprene epoxydiols as 712
precursors to secondary organic aerosol formation: acid-catalyzed reactive uptake studies 713
with authentic compounds, Environ. Sci. Technol., 46, 250-258, 2012. 714
Lin, Y.-H., Zhang, H., Pye, H. O., Zhang, Z., Marth, W. J., Park, S., Arashiro, M., Cui, T., 715
Budisulistiorini, S. H., and Sexton, K. G.: Epoxide as a precursor to secondary organic 716
aerosol formation from isoprene photooxidation in the presence of nitrogen oxides, 717
Proceedings of the National Academy of Sciences, 110, 6718-6723, 2013a. 718
Lin, Y. H., Knipping, E. M., Edgerton, E. S., Shaw, S. L., and Surratt, J. D.: Investigating the 719
influences of SO2 and NH3 levels on isoprene-derived secondary organic aerosol 720
formation using conditional sampling approaches, Atmos. Chem. Phys., 13, 8457-8470, 721
10.5194/acp-13-8457-2013, 2013b. 722
Lin, Y.-H., Budisulistiorini, S. H., Chu, K., Siejack, R. A., Zhang, H., Riva, M., Zhang, Z., Gold, 723
A., Kautzman, K. E., and Surratt, J. D.: Light-absorbing oligomer formation in secondary 724
organic aerosol from reactive uptake of isoprene epoxydiols, Environ. Sci. Technol., 48, 725
12012-12021, 2014. 726
McNeill, V. F., Woo, J. L., Kim, D. D., Schwier, A. N., Wannell, N. J., Sumner, A. J., & Barakat, 727
J. M.: Aqueous-phase secondary organic aerosol and organosulfate formation in 728
atmospheric aerosols: a modeling study, Environ. Sci. Technol., 46(15), 8075-8081, 729
2012. 730
Nenes, A., Pandis, S. N., and Pilinis, C.: ISORROPIA: A new thermodynamic equilibrium model 731
for multiphase multicomponent inorganic aerosols, Aquatic geochemistry, 4, 123-152, 732
1998. 733
Ng, N., Kwan, A., Surratt, J., Chan, A., Chhabra, P., Sorooshian, A., Pye, H., Crounse, J., 734
Wennberg, P., and Flagan, R.: Secondary organic aerosol (SOA) formation from reaction 735
of isoprene with nitrate radicals (NO 3), Atmospheric Chemistry and Physics, 8, 4117-736
4140, 2008. 737
Nguyen, T., Coggon, M., Bates, K., Zhang, X., Schwantes, R., Schilling, K., Loza, C., Flagan, R., 738
Wennberg, P., and Seinfeld, J.: Organic aerosol formation from the reactive uptake of 739
isoprene epoxydiols (IEPOX) onto non-acidified inorganic seeds, Atmospheric 740
Chemistry and Physics, 14, 3497-3510, 2014. 741
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
33
Nguyen, T. B., Bates, K. H., Crounse, J. D., Schwantes, R. H., Zhang, X., Kjaergaard, H. G., 742
Surratt, J. D., Lin, P., Laskin, A., and Seinfeld, J. H.: Mechanism of the hydroxyl radical 743
oxidation of methacryloyl peroxynitrate (MPAN) and its pathway toward secondary 744
organic aerosol formation in the atmosphere, PCCP, 17, 17914-17926, 2015. 745
Nozière, B., Kalberer, M., Claeys, M., Allan, J., D’Anna, B., Decesari, S., Finessi, E., Glasius, 746
M., Grgić, I., and Hamilton, J. F.: The Molecular Identification of Organic Compounds 747
in the Atmosphere: State of the Art and Challenges, Chem. Rev., 10.1021/cr5003485, 748
2015. 749
Pope, C. A., and Dockery, D. W.: Health Effects of Fine Particulate Air Pollution: Lines that 750
Connect, Journal of the Air & Waste Management Association, 56, 709-742, 751
10.1080/10473289.2006.10464485, 2006. 752
Pye, H. O., Pinder, R. W., Piletic, I. R., Xie, Y., Capps, S. L., Lin, Y.-H., Surratt, J. D., Zhang, Z., 753
Gold, A., and Luecken, D. J.: Epoxide pathways improve model predictions of isoprene 754
markers and reveal key role of acidity in aerosol formation, Environ. Sci. Technol., 47, 755
11056-11064, 2013. 756
Riedel, T. P., Lin, Y.-H., Budisulistiorini, S. H., Gaston, C. J., Thornton, J. A., Zhang, Z., Vizuete, 757
W., Gold, A., and Surratt, J. D.: Heterogeneous reactions of isoprene-derived epoxides: 758
reaction probabilities and molar secondary organic aerosol yield estimates, 759
Environmental Science & Technology Letters, 2, 38-42, 2015. 760
Riva, M., Budisulistiorini, S. H., Zhang, Z., Gold, A., and Surratt, J. D.: Chemical characterization 761
of secondary organic aerosol constituents from isoprene ozonolysis in the presence of 762
acidic aerosol, Atmospheric Environment, 2015. 763
Ruthenburg, T. C., Perlin, P. C., Liu, V., McDade, C. E., and Dillner, A. M.: Determination of 764
organic matter and organic matter to organic carbon ratios by infrared spectroscopy with 765
application to selected sites in the IMPROVE network, Atmospheric Environment, 86, 766
47-57, 2014. 767
Schwantes, R. H., Teng, A. P., Nguyen, T. B., Coggon, M. M., Crounse, J. D., St. Clair, J. M., 768
Zhang, X., Schilling, K. A., Seinfeld, J. H., and Wennberg, P. O.: Isoprene NO3 Oxidation 769
Products from the RO2+ HO2 Pathway, The Journal of Physical Chemistry A, 2015. 770
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
34
Simon, H., Bhave, P. V., Swall, J. L., Frank, N. H., and Malm, W. C.: Determining the spatial and 771
seasonal variability in OM/OC ratios across the US using multiple regression, Atmos. 772
Chem. Phys., 11, 2933-2949, 10.5194/acp-11-2933-2011, 2011. 773
Starn, T., Shepson, P., Bertman, S., Riemer, D., Zika, R., and Olszyna, K.: Nighttime isoprene 774
chemistry at an urban-impacted forest site, Journal of Geophysical Research: 775
Atmospheres (1984–2012), 103, 22437-22447, 1998. 776
Stohl, A., Forster, C., Frank, A., Seibert, P., and Wotawa, G.: Technical note: The Lagrangian 777
particle dispersion model FLEXPART version 6.2, Atmos. Chem. Phys., 5, 2461-2474, 778
10.5194/acp-5-2461-2005, 2005. 779
Surratt, J. D., Murphy, S. M., Kroll, J. H., Ng, N. L., Hildebrandt, L., Sorooshian, A., Szmigielski, 780
R., Vermeylen, R., Maenhaut, W., and Claeys, M.: Chemical composition of secondary 781
organic aerosol formed from the photooxidation of isoprene, The Journal of Physical 782
Chemistry A, 110, 9665-9690, 2006. 783
Surratt, J. D., Kroll, J. H., Kleindienst, T. E., Edney, E. O., Claeys, M., Sorooshian, A., Ng, N. L., 784
Offenberg, J. H., Lewandowski, M., and Jaoui, M.: Evidence for organosulfates in 785
secondary organic aerosol, Environ. Sci. Technol., 41, 517-527, 2007a. 786
Surratt, J. D., Lewandowski, M., Offenberg, J. H., Jaoui, M., Kleindienst, T. E., Edney, E. O., and 787
Seinfeld, J. H.: Effect of acidity on secondary organic aerosol formation from isoprene, 788
Environ. Sci. Technol., 41, 5363-5369, 2007b. 789
Surratt, J. D., Chan, A. W., Eddingsaas, N. C., Chan, M., Loza, C. L., Kwan, A. J., Hersey, S. P., 790
Flagan, R. C., Wennberg, P. O., and Seinfeld, J. H.: Reactive intermediates revealed in 791
secondary organic aerosol formation from isoprene, Proceedings of the National 792
Academy of Sciences, 107, 6640-6645, 2010. 793
Tanner, R. L., Bairai, S. T., Olszyna, K. J., Valente, M. L., and Valente, R. J.: Diurnal patterns in 794
PM 2.5 mass and composition at a background, complex terrain site, Atmospheric 795
Environment, 39, 3865-3875, 2005. 796
Wang, W., Kourtchev, I., Graham, B., Cafmeyer, J., Maenhaut, W., and Claeys, M.: 797
Characterization of oxygenated derivatives of isoprene related to 2-methyltetrols in 798
Amazonian aerosols using trimethylsilylation and gas chromatography/ion trap mass 799
spectrometry, Rapid Commun. Mass Spectrom., 19, 1343-1351, 2005. 800
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
35
Xu, L., Guo, H., Boyd, C. M., Klein, M., Bougiatioti, A., Cerully, K. M., Hite, J. R., Isaacman-801
VanWertz, G., Kreisberg, N. M., and Knote, C.: Effects of anthropogenic emissions on 802
aerosol formation from isoprene and monoterpenes in the southeastern United States, 803
Proceedings of the National Academy of Sciences, 112, 37-42, 2015. 804
Zhang, Z., Lin, Y.-H., Zhang, H., Surratt, J., Ball, L., and Gold, A.: Technical Note: Synthesis of 805
isoprene atmospheric oxidation products: isomeric epoxydiols and the rearrangement 806
products cis-and trans-3-methyl-3, 4-dihydroxytetrahydrofuran, Atmospheric Chemistry 807
and Physics, 12, 8529-8535, 2012. 808
809
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
36
Table 1. Sampling schedule during SOAS at the BHM ground site. 810
No. of samples/ day Sampling schedule Dates
2 (regular) Day: 8 am – 7 pm June 1 – June 9
Night: 8 pm – 7 am next day June 13,
June 17 – June 28,
July 2- July 9,
July 15
4 (intensive) Intensive 1: 8 am – 12 pm, June 10 – June 12,
Intensive 2: 1 pm – 3 pm, June 14 – June 16,
Intensive 3: 4 pm – 7 pm, June 29 – June 30,
Intensive 4: 8 pm – 7 am next day July 1,
July 9 – July 14
811
Table 2. Summary of collocated measurements of meteorological variables, gaseous species, and 812
PM2.5 constituents. 813
814 Category Condition Average
SD
Minimum
Maximum
Meteorology Rainfall (in) 0.1
0.2
0.0
1.4
Temp (°C) 26.4
3.0
20.5
32.7
RH (%) 71.5
15.0
36.9
96.1
BP (mbar) 994.2
3.9
984.2
1002.4
SR (W m-2) 303.7
274.5
7.0
885.0
Trace gas (ppbv) O3 31.1
14.8
8.3
62.2
CO 208.7
72.0
99.6
422.9
SO2 0.9
0.8
0.1
3.7
NO 1.3
1.2
0.1
7.0
NO2 6.6
5.1
1.0
22.7
NOx 7.8
6.0
1.3
29.7
NOy 9.1
5.8
2.2
30.4
HNO3 0.3
0.2
0.1
1.0
NH3 1.9
0.8
0.7
4.0
PM2.5 (μg m
-
3
)
OC 7.2
3.2
1.4
14.9
EC 0.6
0.5
0.1
2.7
WSOC 4.0
1.8
0.5
7.5
SO42- 2.0
0.9
0.4
4.9
NO3- 0.1
0.1
0.0
0.8
NH4+ 0.7
0.3
0.2
1.2
Aerosol pH 1.8
0.1
1.6
1.9
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
37
Table 3. Summary of isoprene-derived SOA tracers measured by GC/EI-MS and UPLC/ESI-HR-QTOFMS 815
816
817
818
819
820
821
SOA tracers m/z
Frequency
of detection
(%)a
Max
concentration
(ng/m3)
Mean
concentration
(ng/m3)
Isoprene SOA
Mass fraction
(%)b
% of
total OMc
Measured by GC/EI
-
MS
2
-
methylerythritol
219
9
9.2
1048.
9
26
9.0
33.
8
2.
7
2-methylthreitol 219 100.0
388.9
107.3
13.5 1.1
(E)-2-methylbut-3-ene-1,2,4-triol 231 96.7
878.9
112.7
14.2 1.1
(Z)
-
2
-
methylbut
-
3
-
ene
-
1,2,4
-
triol
231
9
5.8
287.
8
38.9
4.
9
0.
4
2-methylbut-3-ene-1,2,3-triol 231 94.2
503.3
28.9
3.6 0.3
2
-
methylglyceric acid
219
9
3.3
35.
0
10.
8
1.
4
0.
1
cis-3-MeTHF-3,4-diol 262 22.5
98.9
6.9
0.9 0.1
trans-3-MeTHF-3,4-diol 262 10.0
137.6
8.6
1.1 0.1
IEPOX
-
derived dimer
333
1
0
.0
2.
2
0.
0
0.
0
0.
0
Levoglucosan 204 100.0
922.6
98.7
- 1.0
Measured by UPLC/ESI-HR-
QTOFMS
IEPOX-derived OSs
C
5
H
11
O
7
S
-
215 100.0
864.9
164.5
20.7 1.6
C
10
H
21
O
10
S
-
333
1
.7
0.
3
0.0
0.0
0.
0
MAE-derived OS
C
4
H
7
O
7
S
-
199 100.0
35.7
7.2
1.9 0.1
GA sulfate
C
2
H
3
O
6
S
-
155 100.0
75.2
26.2
3.3 0.3
Methylglyoxal-derived OS
C
3
H
5
O
6
S
-
169 97.5
10.5
2.7
0.3 0.0
Isoprene-derived OSs
C
5
H
7
O
7
S
-
211 97.5
5.2
1.4
0.2 0.0
C
5
H
10
NO
9
S
-
260
9
0
.0
3.
9
0.
3
0.0
0.
0
C
5
H
9
N
2
O
11
S
-
305 5.0
3.3
2.9
0.4 0.0
Hydroxyacetone-derived OS
C
2
H
3
O
5
S
-
139 30.8
2.6
0.2
0.0 0.0
a
Total filters = 120
b Mass fraction is the contribution of each species among total known isoprene-derived SOA mass detected by GC/EI
MS and UPLC/ESI-HR-QTOFMS
c OM/OC = 1.6
d
d
d
d
d
d
d
d
d
OA tracers quantified by authentic standards
e SOA tracers quantified by 2-methyltetrols as a surrogate standard
f SOA tracer quantified by IEPOX-derived OS (m/z 215) as a surrogate standard
g SOA tracers quantified by propyl sulfate as a surrogate standard
d
e
e
e
e
f
g
g
g
m/z
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
38
Table 4. Overall correlation (r2) of isoprene-derived SOA tracers and collocated measurements at 822
BHM during 2013 SOAS campaign. 823
SOA tracers CO O 3 NOx NOy SO2 NH3 SO4 NO3 NH4 OC WSOC pH
MAE/HMML-derived SOA
tracers 0.07 0.26 0.00 0.01 0.06 0.11 0.33 0.01 0 .18 0.47 0.20 0.00
2-methylglyceric acid 0.01 0.26 0.01 0.00 0.01 0.07 0.10 0.00 0.06 0.19 0.02 0.00
MAE-derived OS 0.10 0.14 0.00 0.02 0.07 0.09 0.38 0.01 0.18 0.32 0.23 0.01
IEPOX-derived SOA
tracers
0.04 0.05 0.00 0.01 0.05 0.01 0.36 0.00 0.21 0.24 0.12 0.00
2-methylerythritol 0.00 0.16 0.03 0.02 0.01 0.00 0.30 0.02 0.1 8 0.18 0.19 0.00
2-methylthreitol 0.00 0.13 0.02 0.03 0.02 0.00 0.20 0.01 0.16 0.17 0.15 0.00
(E)-2-methylbut-3-ene-1,2,4-triol 0.07 0.00 0.02 0 .01 0.07 0.00 0.15 0.00 0.19 0.11 0.04 0.00
(Z)-2-methylbut-3-ene-1,2,4-triol 0.04 0.00 0.00 0.00 0.06 0.00 0.28 0.00 0.20 0.04 0.00 0.00
2-methylbut-3-ene-1,2,3-triol 0.02 0.00 0.03 0.00 0.00 0.02 0.32 0.01 0.03 0.17 0.04 0.00
IEPOX-derived OS 0.02 0 .14 0.03 0.00 0.00 0.00 0.27 0.00 0.16 0.29 0.29 0.00
IEPOX dimer 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Other isoprene SOA tracers
GA sulfate
C2H3O6S
-
0.30 0.23 0.01 0.00 0.08 0.09 0.27 0.00 0.19 0.38 0.18 0.00
Methylglyoxal-derived OS
C3H5O6S
-
0.14 0.04 0.02 0.03 0.03 0.07 0.31 0.02 0 .25 0.21 0.24 0.00
Isoprene-derived OSs
C5H7O7S
-
0.01 0.23 0.03 0.01 0.00 0.02 0.21 0.00 0.16 0.31 013 0.00
C5H10NO9S
-
0.17 0.00 0.12 0.14 0.10 0.14 0.31 0.16 0.23 0.20 0.07 0.00
C5H9N2O11S
-
*
0.32 0.71 0.66 0.58 0.42 0.02 0.68 0.50 0.42 0.00 0.50 0.00
Hydroxyacetone-derived OS
C2H3O5S
-
0.02 0.10 0.08 0.07 0.05 0.00 0.00 0.03 0.00 0.01 0.01 0.00
Other tracer
Levoglucosan 0.00 0.09 0.02 0 .01 0.02 0 .00 0.00 0.02 0 .00 0.08 0.04 0.01
* Found only in 6 of 120 filters 824
The correlations in this table are positive. 825
826
827
828
829
830
831
832
833
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
39
Table 5. Summary of isoprene-derived SOA tracers from the three SOAS ground sites: BHM, 834
CTR, and LRK. 835
836
SOA tracers
Urban Rural
BHM CTR LRK
Mean
(ng m-3)
Average
amount
detected
tracers
(%)
Mean
(ng m-3)
Average
amount
detected
tracers
(%)
Mean
(ng m-3)
Average
amount
detected
tracers
(%)
MAE/HMML derived SOA
MAE/HMML
-
derived
OS
7.2
1.1
10.2
1.3
8.2
.1 8
2
-
methylglyceric acid
10.4
1.7
5.1
0.7
7.5
.1 6
IEPOX derived SOA
IEPOX
-
derived
OS
164.5
24.3
207.1
26.8
139.2
.30 3
IEPOX
-
derived dimer
OS
0.04
0.00
0.7
0.1
1.1
.0 2
2
-
methylerythritol
266.7
37.9
204.8
26.5
120.7
.26 3
2
-
methylthreitol
107.3
15.8
73.7
9.5
42.4
.9 2
(E)
-
2
-
methylbut
-
3
-
ene
-
1,2,4
-
triol
109.0
12.3
137.3
17.8
98.8
.21 5
(Z)
-
2
-
methylbut
-
3
-
ene
-
1,2,4
-
triol
37.3
4.1
50.7
6.6
29.1
.6 1
2
-
methylbut
-
3
-
ene
-
1,2,3
-
triol
23.4
2.5
26.1
3.4
16.5
.3 6
trans
-
3
-
MeTHF
-
3,4
-
diol
8.6
1.0
0.0
0.0
2.7
.0 6
cis
-
3
-
MeTHF
-
3,4
-
diol
6.8
1.0
0.2
0.0
1.7
.0 4
837
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
40
838
839
840
841
842
843
844
845
846
847
848
849
Figure 1. Wind rose illustrating wind direction during the campaign at the BHM site. Bars indicate 850
direction of incoming wind, with 0 degrees set to geographic north. Length of bar size indicates 851
frequency with color segments indicating the wind speed in m s-1. 852
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
41
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
Figure 2. Time series of (a) meteorological data, (b) trace gases, (c) PM2.5 constituents, (d) 874
MAE/HMML-derived SOA tracers and (e) IEPOX-derived SOA tracers during the 2013 SOAS 875
campaign at the BHM site. 876
2000
1500
1000
500
0
Mass conc. (ng m-3)
11/06/2013 21/06/2013 01/07/2013 11/07/2013
Date and Time (Local)
100
80
60
40
20
0
Mass conc. (ng m-3)
90
80
70
60
50
40
RH (%)
32
28
24
Temp (C)
4
3
2
1
0
Rainfall (inch)
50
40
30
20
10
0
SO2, NOx, NH3 (ppb)
400
200
0
-200
-400
CO (ppb)
120
80
40
0
O3 (ppb)
Total MAE/HMML-SOA trace rs
(a)
Meteor
ology
(b) Trace gas
(c) PM
2.5
composition
(d) MAE/HMML-derived SOA
(e) IEPOX-derived SOA
RH, Temperature, Rainfall
CO
O3, NOx, SO2, NH3
OC, WSOC
SO4, NH4, NO3
2-methylglyceric aid (2-MG)
MAE-derived organosulfate (MAE-OS)
2-methylthrietol
2-methylerythritol
(z)-2-methylbut-3-ene-1,2,4,triol
2-methylbut-3-ene-1,2,3-triol
(E)-2-methylbut-3-ene-1,2,4-triol
IEPOX-derived organosulfate
Total IEPOX-SOA tracer s
10
8
6
4
2
0
SO4, NO3, NH4 (ug m-3
)
-20
-10
0
10
20
OC, WSOC (ug m-3)
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
42
877
878
879
880
881
882
(a) (b) (c) 883
884
885
886
887
888
889
890
(d) (e) (f) 891
892
Figure 3. Correlation of MAE-derived SOA tracers with (a) daytime NO2, (b) daytime O3, (c) 893
daytime P[NO3], (d) nighttime NO2, (e) nighttime O3, and (f) nighttime P[NO3]. Nighttime P[NO3] 894
correlation suggests that NO3 radical chemistry could explain some fraction of the MAE/HMML-895
derived SOA tracer concentrations. 896
897
898
5x106
4321
Daytime P[NO3]
r2 = 0.07
6050403020
Daytime O3 (ppb)
r2 = 0.48
100
80
60
40
20
0
Mass conc. (ng m-3)
8642
Daytime NO2 (ppb)
r2 = 0.04
100
80
60
40
20
0
Mass conc. (ng m-3)
2015105
Nighttime NO2 (ppb)
r2 = 0.26
6050403020100
Nighttime O3 (ppb)
r2 = 0.08
6x106
5432
Nighttime P[NO3]
r2 = 0.57
Daytime NO
2
(ppb)
D
aytime
O
3
(ppb)
D
aytime P[NO
3
]
Nighttime
NO
2
(ppb)
Nighttime
O
3
(ppb)
Nighttime P[NO
3
]
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
43
899
900
901
902
903
904
(a) (b) (c) 905
906
907
908
909
910
911
(d) (e) (f) 912
Figure 4. Correlation of IEPOX-derived SOA tracers with (a) daytime NO2, (b) daytime O3 , (c) 913
daytime P[NO3], (d) nighttime NO2, (e) nighttime O3, and (f) nighttime P[NO3]. Nighttime P[NO3] 914
correlation suggests that NO3 radical chemistry could explain some fraction of the IEPOX-derived 915
SOA tracer concentrations. 916
2000
1500
1000
500
0
Mass conc. (ng m-3)
8642
r2 = 0.00
6050403020
r2 = 0.10
40353025201510
r2 = 0.09
2000
1500
1000
500
0
Mass conc. (ng m-3)
2015105
r2 = 0.04
6x10
6
5432
r2 = 0.26
5x10
6
4321
r2 = 0.06
Daytime NO
2
(ppb)
Daytime O
3
(ppb)
Daytime P[NO
3
]
Nighttime NO
2
(ppb)
Nighttime O
3
(ppb)
Nighttime P[NO
3
]
Atmos. Chem. Phys. Discuss., doi:10.5194/acp-2015-983, 2016
Manuscript under review for journal Atmos. Chem. Phys.
Published: 19 January 2016
c
Author(s) 2016. CC-BY 3.0 License.
... Many recent publications report instead chemical shift data about secondary organic compounds produced by terpene oxidation encompassing numerous key tracers such as methyltetrols, 2-methylglyceric acid, isoprene-epoxydiol (IEPOX) -derived compounds (e.g., [43], [44], [32]), along with a various range of carboxylic and oxo-carboxylic acids and their dimers formed by the oxidation of monoterpenes (-pinene, carene, limonene) ( [45], [34], [37], [33]). Additional chemical shift data were let available for some of the above species functionalized with hydroperoxyl-, nitrate-or sulfate groups ( [46], [44], [47], [35], [48], [49], [50], [36]). There is then a paucity of NMR data for anthropogenic secondary organic compounds with the exception of phenolic compounds. ...
... The spreading of NMR applications in aerosol studies claims for strengthening the network of a growing NMR atmospheric community for better standardization of protocols for the acquisition and processing of the spectra, as well as for sharing chemical shift data and libraries of reference NMR spectra. Methyl-tetrols, isoprene epoxydiols (IEPOX), methylthreonic, methylthreonic and methylglyceric acids, others [31], [43] [44], [47], [32], [41], [106], [107], [50], [108] ...
... 25,31,52−59 Consequently, uncertainty is introduced to the evaluation of the importance of individual OSs as well as OS as a compound group. A few studies 30,34,36,58,60−66 used authentic synthetic standards to quantify individual OSs derived from isoprene 30,34,64 and small OSs containing two or three carbon atoms. 36,61,63−65 Apparently, the lack of OS authentic standards becomes a hurdle for accurate quantification and comprehensive formation mechanism for OS compounds. ...
... 25,31,52−59 Consequently, uncertainty is introduced to the evaluation of the importance of individual OSs as well as OS as a compound group. A few studies 30,34,36,58,60−66 used authentic synthetic standards to quantify individual OSs derived from isoprene 30,34,64 and small OSs containing two or three carbon atoms. 36,61,63−65 Apparently, the lack of OS authentic standards becomes a hurdle for accurate quantification and comprehensive formation mechanism for OS compounds. ...
Article
Organosulfates (OSs) derived from biogenic volatile organic compounds are important compounds signifying interactions between anthropogenic sulfur pollution and natural emissions. In this work, we substantially expand the OS standard library through the chemical synthesis of 26 -hydroxy OS standards from eight monoterpenes (i.e., - & -pinene, limonene, sabinene, 3-carene, terpinolene,-& -terpinene) and two sesquiterpenes (i.e., -humulene and -caryophyllene). The sulfation of unsymmetrically substituted 1,2-diol intermediates produced a regioisomeric mixture of two OSs. The major regioisomeric OSs were isolated and purified for full NMR characterization while the minor regioisomers could only be determined by liquid chromatograph-mass spectroscopy. The tandem mass spectra of the molecular ion formed through electrospray ionization confirmed the formation of abundant bisulfate ion fragment (m/z 97) and certain minor ion fragments characteristic of the carbon backbone. Knowledge of the MS/MS spectra and chromatographic retention times for authentic standards allows us to identify -hydroxy OSs derived from six monoterpenes and -caryophyllene in ambient samples. Notably, among two possible regioisomers of -hydroxy OSs, we only detected the isomers with the sulfate group at the less substituted carbon position derived from -pinene, limonene, sabinene, 3-carene, and terpinolene in the ambient samples. This observation sheds light on the atmospheric OS formation mechanisms.
... Similarly, elevated PM 2.5 observed in New York and Wisconsin has been attributed to Ohio River Valley emissions. Transported pollutants can impact biogenic secondary organic aerosol (SOA) formation in remote locations(Carlton et al., 2010; Emanuelsson et al., 2013;Rattanavaraha et al., 2016; Xu et al., 2015).Finally, prior and on-going studies through the IMPROVE program in rural locations throughout North America have investigated both transported and local contributions to the aerosol populations (Hand et al., 2011). Uncertainty in the contributions of long-range aerosols and limited measurements in remote areas can lead to inaccuracies in modeling of aerosol source contributions. ...
Thesis
Atmospheric aerosols have significant impacts on air quality, climate, and human health, yet analytical and logistical challenges have limited our ability to measure these aerosol particles, particularly in remote regions. In this dissertation, individual atmospheric particles were chemically characterized in rural northern Michigan and remote northern Alaska for the first time. To enable this measurements, Chapter 2 details the construction and characterization of an updated aircraft-capable aerosol time-of-flight mass spectrometer (A-ATOFMS), capable of measuring size-resolved chemical composition of 0.07 – 1.6 µm individual particles up to 40 Hz with lower mass (~25 kg saved) and power (~600 W saved) consumption than the previous A-ATOFMS. Chapter 3 discusses size-resolved chemical composition of atmospheric aerosols in northern Michigan while the site was influenced by Canadian wildfire, urban, and local forest air masses. Throughout the study, long-range transported biomass burning aerosols were the cores of particles primarily consisting, by mass, of secondary organic aerosol from the oxidation of volatile organic compounds emitted from both wildfires and forests. In Chapter 4, we identified 14 periods of ultrafine particle growth at the same field site. Urban air mass influence during the daytime led to the highest observed growth rates, likely due to increased atmospheric oxidant levels producing condensable material. Nighttime wildfire air masses were likely influenced by increased SO2 and NO2 in the plumes leading to NO3 radical oxidation. TEM-EDX showed contributions from sulfur, carbon, and oxygen down to 20 nm particles, suggesting contributions from H2SO4 and SOA. As particle growth was previously thought to be suppressed in this isoprene-rich forest, these measurements represent a source of particles not previously considered in this environment. Chapters 5 – 6 discuss the results from field campaigns conducted in the Alaskan Arctic. In Chapter 5, I show results of A-ATOFMS and scanning electron microscopy with energy-dispersive x-ray spectroscopy (SEM-EDX) analyses of atmospheric particles transported to Utqiaġvik, AK from the Prudhoe Bay oil fields, located hundreds of kilometers to the east, in comparison to the pristine Arctic Ocean background. During Arctic Ocean influence, fresh sea spray aerosol (SSA) was the primary contributor to aerosol number concentrations, compared to transported organic carbon and aged SSA particles during Prudhoe Bay air masses. Chapter 6 details the 2016 field campaign within the Prudhoe Bay oil fields, where we deployed the A-ATOFMS to characterize local oil field combustion plumes and the overall oil field background aerosol population; these were the first single particle measurements within an Arctic oil field. Diesel and natural gas combustion were the major influences on the aerosol population, with unique amine-containing particles identified from the processing of natural gas. Overall, the results from these field campaigns, aided by the newly constructed A-ATOFMS, provided new insights into the chemical composition of local and transported atmospheric particles on rural and remote environments influenced by the changing climate.
... Studies about isoprene oxidation in the 1990's suggested that this species does not contribute significantly to the atmospheric SOA budget (Pandis et al., 1991) because early-generation oxidation species from isoprene are highly volatile. Subsequent laboratory chamber experiments and field observations over the past 20 years have 10 shown that multi-generation oxidation of isoprene and its oxidation products contribute considerably to organic aerosol mass in the atmosphere by the formation of SVOCs, LVOCs and extremely low-volatility organic compounds (ELVOCs) (Claeys et al., 2004b;Dommen et al., 2009;Edney et al., 2005;Kroll et al., 2005;Kleindienst et al., 2006;Kroll et al., 2006;Kleindienst et al., 2007;Matsunaga et al., 2005;Ng et al., 2008;Nguyen et al., 2010;Surratt et al., 2010;Lin et al., 2011;Mao et al., 2013;Hu et al., 2015;Jokinen et al., 2015;Krechmer et al., 2015;Song et al., 2015;Xu et al., 2015;Xiong et al., 2015;Kourtchev 15 et al., 2016;Lopez-Hilfiker et al., 2016;Rattanavaraha et al., 2016;Riva et al., 2016). Several chemical pathways have been put forward to describe the formation of semi-volatile and low-volatility higher-generation oxidation products via the oxidation of early-generation compounds (Bates et al., 2014;Kameel et al., 2013;Kroll et al., 2006;Mao et al., 2013;Surratt et al., 2010;Worton et al., 2013). ...
Preprint
Full-text available
This study presents a characterization of the hygroscopic growth behaviour and effects of different inorganic seed particles on the formation of secondary organic aerosols (SOA) from the dark ozone-initiated oxidation of isoprene at low NOx conditions. We performed simulations of isoprene oxidation using a gas-phase chemical reaction mechanism based on the Master Chemical Mechanism (MCM) in combination with an equilibrium gas–particle partitioning model to predict the SOA concentration. The equilibrium model accounts for non-ideal mixing in liquid phases, including liquid–liquid phase separation (LLPS), and is based on the AIOMFAC model for mixture non-ideality and the EVAPORATION model for pure compound vapour pressures. Measurements from the Cosmics Leaving Outdoor Droplets (CLOUD) chamber experiments conducted at the European Organization for Nuclear Research (CERN) for isoprene ozonolysis cases, were used to aid in parameterizing the SOA yields at different atmospherically relevant temperatures, relative humidity (RH) and reacted isoprene concentrations. To represent the isoprene ozonolysis-derived SOA, a selection of organic surrogate species is introduced in the coupled modelling system. The model predicts a single, homogeneously mixed particle phase at all relative humidity levels for SOA formation in the absence of any inorganic seed particles. In the presence of aqueous sulfuric acid or ammonium bisulfate seed particles, the model predicts LLPS to occur below ~80 % RH, where the particles consist of an inorganic-rich liquid phase and an organic-rich liquid phase; however, with significant amounts of bisulfate and water partitioned to the organic-rich phase. The measurements show an enhancement in the SOA amounts at 85 % RH compared to 35 % RH for both the seed-free and seeded cases. The model predictions of RH-dependent SOA yield enhancements at 85 % RH vs. 35 % RH are 1.80 for a seed-free case, 1.52 for the case with ammonium bisulfate seed and 1.06 for the case with sulfuric acid seed. Predicted SOA yields are enhanced in the presence of an aqueous inorganic seed, regardless of the seed type (ammonium sulfate, ammonium bisulfate or sulfuric acid) in comparison with seed-free conditions at the same RH level. We discuss the comparison of model-predicted SOA yields with a selection of other laboratory studies on isoprene SOA formation conducted at different temperatures and for a variety of reacted isoprene concentrations.
... In summary, in areas such as the southeastern United States, where isoprene emissions are high, understanding the specific cellular responses to isoprene-derived SOA is crucial to public health. 42,43,54,57 This study provides evidence that miRNAs play a role in isoprenederived SOA-induced morbidity and do so in a manner that is specific to chemical composition of the SOA. ...
Article
Exposure to fine particulate matter (PM2.5), of which secondary organic aerosol (SOA) is a major constituent, is linked to adverse health outcomes including cardiovascular disease, lung cancer and preterm birth. Atmospheric oxidation of isoprene, the most abundant non-methane hydrocarbon emitted into Earth’s atmosphere primarily from vegetation, contributes to SOA formation. Isoprene-derived SOA has previously been found to alter inflammatory/oxidative stress genes. Micro RNAs (miRNAs) are epigenetic regulators that serve as post-transcriptional modifiers and key mediators of gene expression. To assess whether isoprene-derived SOA alters miRNA expression, BEAS-2B lung cells were exposed to laboratory-generated isoprene-derived SOA constituents derived from the acid-driven multiphase chemistry of authentic methacrylic acid epoxide (MAE) or isomeric isoprene epoxydiols (IEPOX) with acidic sulfate aerosol particles. These IEPOX- and MAE-derived SOA constituents have been shown to be measured in large quantities within PM2.5 collected from isoprene-rich areas affected by acidic sulfate aerosol particles derived from human activities. A total of 29 miRNAs were identified as differentially expressed when exposed to IEPOX-derived SOA and two when exposed to MAE-derived SOA, a number of which are inflammatory/oxidative stress associated. These results suggest that miRNAs may modulate the inflammatory/oxidative stress response to SOA exposure thereby advancing the understanding of airway cell epigenetic response to SOA.
... OS mass ranges from 400 to 1500 ng m −3 , at the high end of previous studies (Tables S2−S4). 26,27,32,80,81 While the sum of OS in the SE-U.S. is significantly different from that of downwind Manaus, isoprene-OS nevertheless represents the predominant OS in both areas. During the 2013 SOAS campaign, total mass concentrations of organosulfur compounds were also determined by IR-ICP-MS ( Figure S4). ...
Article
Acid-driven multiphase chemistry of isoprene epoxydiols (IEPOX), key isoprene oxidation products, with inorganic sulfate aerosol yields substantial amounts of secondary organic aerosol (SOA) through the formation of organosulfur compounds. The extent and implications of inorganic-to-organic sulfate conversion, however, are unknown. In this article, we demonstrate that extensive consumption of inorganic sulfate occurs, which increases with the IEPOX-to-inorganic sulfate concentration ratio (IEPOX/Sulfinorg), as determined by laboratory measurements. Characterization of the total sulfur aerosol observed at Look Rock, Tennessee, from 2007 to 2016 shows that organosulfur mass fractions will likely continue to increase with ongoing declines in anthropogenic Sulfinorg, consistent with our laboratory findings. We further demonstrate that organosulfur compounds greatly modify critical aerosol properties, such as acidity, morphology, viscosity, and phase state. These new mechanistic insights demonstrate that changes in SO2 emissions, especially in isoprene-dominated environments, will significantly alter biogenic SOA physicochemical properties. Consequently, IEPOX/Sulfinorg will play an important role in understanding the historical climate and determining future impacts of biogenic SOA on the global climate and air quality.
... To some extent, these strongly elevated concentrations might be connected to the urban sampling location. This was further supported by measurements at rural sampling locations in the same region, which gave somewhat lower OS concentrations ( Hettiyadura et al. 2017;Rattanavaraha et al. 2016). Nonetheless, concentrations for single isoprene-derived OSs were still reported to be in the range of 0.2 to 668 ng m À3 , and thus, larger than most MT-OS concentrations determined in our study. ...
Article
Environmental contextSecondary organic aerosols account for a major fraction of atmospheric particulate matter, affecting both climate and human health. Organosulfates, abundant compounds in organic aerosols, are difficult to measure because of the lack of authentic standards. Here we quantify terpene-derived organosulfates in atmospheric particulate matter at a rural site in Germany and at the North China Plain using a combined target/non-target high-resolution mass spectrometry approach. AbstractOrganosulfates (OSs) are a ubiquitous class of compounds in atmospheric aerosol particles. However, a detailed quantification of OSs is commonly hampered because of missing authentic standards and the abundance of unknown OSs. Using a combined targeted and untargeted approach of high-resolution liquid chromatography–Orbitrap mass spectrometry (LC–Orbitrap MS), we quantified for the first time the total concentrations of known and unknown monoterpene (MT) and sesquiterpene (SQT) OSs in summertime PM10 particulate matter from field studies in rural Germany (MEL) and the North China Plain (NCP). At each site, we observed more than 50 MT-OSs, 13 of which were detectable at both sites. For both locations, median concentrations of MT-OSs were in the range of 10 to 40ngm−3, to which the 13 common MT-OSs contributed on average >50%. The main contributor to MT-OSs was C9H16O7S (MT-OS 267) with average mass concentrations of 2.23 and 6.38ngm−3 for MEL and NCP respectively. The concentrations of MT-OSs correlated with the concentrations of MT oxidation products only for MEL. For NCP, the low concentrations of MT oxidation products (i.e. typically <1ngm−3) suggested a suppression of carboxylic acid formation under high concentrations of NOx and particulate sulfate. Furthermore, we observed 17 SQT-OSs for the MEL samples, whereas 40 SQT-OSs were detected in the NCP samples. Only five of these SQT-OSs were detectable at both sites. Correspondingly, the total concentrations of SQT-OSs were larger for NCP than for MEL, which suggested large differences in the particle chemistry. In particular, aerosol acidity was found to be a key factor during SQT-OS formation, and was probably not sufficient in the PM10 from MEL.
Chapter
Full-text available
The effects of atmospheric fine particles on human health have become an utmost concern worldwide. Particulate matter is a complex and dynamic combination of a mixture of solid and liquid substances with several biological and chemical components. Various toxicological and epidemiological studies indicated that the fine particles create several health issues such as respiratory and cardiopulmonary disorders. The present chapter provides the information regarding regulations and standards set by various countries and organizations to regulate the atmospheric concentration of fine particles and discuss the primary and secondary sources of fine particulate pollution. This chapter demonstrated the biological and chemical components of fine particles that play a critical role in the toxicological implications of fine particulates. In addition, the justifications for the origin or sources of biological and chemical compositions and their impacts on human health become a concern in this chapter. The current chapter also aims to provide a brief overview of the molecular mechanisms connecting fine particulate exposure and health effects.KeywordsFine particlesSourcesCompositionBio-aerosolToxicityHealth effectsMechanism of action
Article
Anthropogenic emissions alter biogenic secondary organic aerosol (SOA) formation from naturally emitted volatile organic compounds (BVOCs). We review the major laboratory and field findings with regard to effects of anthropogenic pollutants (NOx, anthropogenic aerosols, SO2, NH3) on biogenic SOA formation. NOx participate in BVOC oxidation through changing the radical chemistry and oxidation capacity, leading to a complex SOA composition and yield sensitivity towards NOx level for different or even specific hydrocarbon precursors. Anthropogenic aerosols act as an important intermedium for gas—particle partitioning and particle-phase reactions, processes of which are influenced by the particle phase state, acidity, water content and thus associated with biogenic SOA mass accumulation. SO2 modifies biogenic SOA formation mainly through sulfuric acid formation and accompanies new particle formation and acid-catalyzed heterogeneous reactions. Some new SO2-involved mechanisms for organosulfate formation have also been proposed. NH3/amines, as the most prevalent base species in the atmosphere, influence biogenic SOA composition and modify the optical properties of SOA. The response of SOA formation behavior to these anthropogenic pollutants varies among different BVOCs precursors. Investigations on anthropogenic—biogenic interactions in some areas of China that are simultaneously influenced by anthropogenic and biogenic emissions are summarized. Based on this review, some recommendations are made for a more accurate assessment of controllable biogenic SOA formation and its contribution to the total SOA budget. This study also highlights the importance of controlling anthropogenic pollutant emissions with effective pollutant mitigation policies to reduce regional and global biogenic SOA formation.
Article
Full-text available
Organosulfates (OSs) have recently been observed to be a potentially important constituent of secondary organic aerosol (SOA); however, their molecular characterization in highly polluted atmospheres has not been probed in detail. This study thoroughly presents the characterization of OSs in polluted air and demonstrates their seasonal and diurnal variations, formation mechanisms, and contributions to organic aerosol. Atmospheric PM2.5 samples were collected from an urban Shanghai site across the winter and summer of 2017. OSs were characterized by ultra‐high‐performance liquid chromatography (UHPLC) coupled with Orbitrap mass spectrometry (MS). Based on exact mass formulae in conjunction with previous chamber studies, hundreds of sulfur‐containing compounds were tentatively identified as OSs. The number and abundance of OSs increased significantly during pollution episodes. The OSs in the clean aerosol samples were dominant in biogenic products, whereas the OSs in the polluted winter samples had distinctive anthropogenic characteristics. Aromatics and long‐chain alkanes from anthropogenic emissions might be their precursors. By using synthesized standards, the total concentrations of 14 quantified OSs ranged 21.6–161 ng m⁻³ in summer and 5.85–84.3 ng m⁻³ in winter, respectively. Among these OSs, glycolic acid sulfate was the most abundant species (1.13–122 ng m⁻³). Further analysis of their seasonal and diurnal variations suggests possible contributions from multiple formation mechanisms, including acid‐catalyzed and NO3‐initiated oxidation reactions. Our results highlight that increased anthropogenic pollutant emissions (e.g., NOx and SO2) can significantly enhance the SOA burden in biogenically influenced urban areas.
Article
Full-text available
Fast measurements of aerosol and gas-phase constituents coupled with the ISORROPIA-II thermodynamic equilibrium model are used to study the partitioning of semivolatile inorganic species and phase state of Mexico City aerosol sampled at the T1 site during the MILAGRO 2006 campaign. Overall, predicted semivolatile partitioning agrees well with measurements. PM2.5 is insensitive to changes in ammonia but is to acidic semivolatile species. Semi-volatile partitioning equilibrates on a timescale between 6 and 20 min. When the aerosol sulfate-to-nitrate molar ratio is less than 1, predictions improve substantially if the aerosol is assumed to follow the deliquescent phase diagram. Treating crustal species as "equivalent sodium" (rather than explicitly) in the thermodynamic equilibrium calculations introduces important biases in predicted aerosol water uptake, nitrate and ammonium; neglecting crustals further increases errors dramatically. This suggests that explicitly considering crustals in the thermodynamic calculations are required to accurately predict the partitioning and phase state of aerosols.
Article
Full-text available
Isoprene epoxydiol (IEPOX) isomers are key gas-phase intermediates of isoprene atmospheric oxidation. Secondary organic aerosols derived from such intermediates have important impacts on air quality and health. We report here convergent and unambiguous pathways developed for the synthesis of isomeric IEPOX species and the rearrangement products cis- and trans-3-methyl-3,4-dihydroxytetrahydrofuran in good yield. The availability of such compounds is necessary to expedite research on isoprene atmospheric oxidation mechanisms and subsequent aerosol formation as well as the toxicological properties of the aerosols.
Article
Full-text available
Substantial amounts of secondary organic aerosol (SOA) can be formed from isoprene epoxydiols (IEPOX), which are oxidation products of isoprene mainly under low-NO conditions. Total IEPOX-SOA, which may include SOA formed from other parallel isoprene oxidation pathways, was quantified by applying positive matrix factorization (PMF) to aerosol mass spectrometer (AMS) measurements. The IEPOX-SOA fractions of organic aerosol (OA) in multiple field studies across several continents are summarized here and show consistent patterns with the concentration of gas-phase IEPOX simulated by the GEOS-Chem chemical transport model. During the Southern Oxidant and Aerosol Study (SOAS), 78 % of PMF-resolved IEPOX-SOA is accounted by the measured IEPOX-SOA molecular tracers (2-methyltetrols, C5-Triols, and IEPOX-derived organosulfate and its dimers), making it the highest level of molecular identification of an ambient SOA component to our knowledge. An enhanced signal at C5H6O+ (m/z 82) is found in PMF-resolved IEPOX-SOA spectra. To investigate the suitability of this ion as a tracer for IEPOX-SOA, we examine fC5H6O (fC5H6O= C5H6O+/OA) across multiple field, chamber, and source data sets. A background of ~ 1.7 ± 0.1 ‰ (‰ = parts per thousand) is observed in studies strongly influenced by urban, biomass-burning, and other anthropogenic primary organic aerosol (POA). Higher background values of 3.1 ± 0.6 ‰ are found in studies strongly influenced by monoterpene emissions. The average laboratory monoterpene SOA value (5.5 ± 2.0 ‰) is 4 times lower than the average for IEPOX-SOA (22 ± 7 ‰), which leaves some room to separate both contributions to OA. Locations strongly influenced by isoprene emissions under low-NO levels had higher fC5H6O (~ 6.5 ± 2.2 ‰ on average) than other sites, consistent with the expected IEPOX-SOA formation in those studies. fC5H6O in IEPOX-SOA is always elevated (12–40 ‰) but varies substantially between locations, which is shown to reflect large variations in its detailed molecular composition. The low fC5H6O (< 3 ‰) reported in non-IEPOX-derived isoprene-SOA from chamber studies indicates that this tracer ion is specifically enhanced from IEPOX-SOA, and is not a tracer for all SOA from isoprene. We introduce a graphical diagnostic to study the presence and aging of IEPOX-SOA as a triangle plot of fCO2 vs. fC5H6O. Finally, we develop a simplified method to estimate ambient IEPOX-SOA mass concentrations, which is shown to perform well compared to the full PMF method. The uncertainty of the tracer method is up to a factor of ~ 2, if the fC5H6O of the local IEPOX-SOA is not available. When only unit mass-resolution data are available, as with the aerosol chemical speciation monitor (ACSM), all methods may perform less well because of increased interferences from other ions at m/z 82. This study clarifies the strengths and limitations of the different AMS methods for detection of IEPOX-SOA and will enable improved characterization of this OA component.
Article
Full-text available
Data from the Interagency Monitoring of Protected Visual Environments (IMPROVE) network are used to estimate organic mass to organic carbon (OM/OC) ratios across the United States by extending previously published multiple regression techniques. Our new methodology addresses common pitfalls of multiple regression including measurement uncertainty, colinearity of covariates, and dataset selection. As expected, summertime OM/OC ratios are larger than wintertime values across the US with all regional median OM/OC values tightly confined between 1.8 and 1.95. Further, we find that OM/OC ratios during the winter are distinctly larger in the eastern US than in the West (regional medians are 1.58, 1.64, and 1.85 in the great lakes, southeast, and northeast regions, versus 1.29 and 1.32 in the western and central states). We find less spatial variability in long-term averaged OM/OC ratios across the US (90% of our multiyear regressions predicted OM/OC ratios between 1.37 and 1.94) than previous studies (90% of OM/OC estimates from a previous regression study fell between 1.30 and 2.10). We attribute this difference largely to the inclusion of EC as a covariate in previous regression studies. Due to the colinearity of EC and OC, we believe that up to one-quarter of the OM/OC estimates in a previous study are biased low. In addition to estimating OM/OC ratios, our technique reveals trends that may be contrasted with conventional assumptions regarding nitrate, sulfate, and soil across the IMPROVE network. For example, our regressions show pronounced seasonal and spatial variability in both nitrate volatilization and sulfate neutralization and hydration.
Article
Full-text available
Long-term (1999 to 2013) data from the Southeastern Aerosol Research and Characterization (SEARCH) network are used to show that anthropogenic emission reductions led to important decreases in fine-particle organic aerosol (OA) concentrations in the southeastern US On average, 45 % (range 25 to 63 %) of the 1999 to 2013 mean organic carbon (OC) concentrations are attributed to combustion processes, including fossil fuel use and biomass burning, through associations of measured OC with combustion products such as elemental carbon (EC), carbon monoxide (CO), and nitrogen oxides (NOx). The 2013 mean combustion-derived OC concentrations were 0.5 to 1.4 µg m−3 at the five sites operating in that year. Mean annual combustion-derived OC concentrations declined from 3.8 ± 0.2 µg m−3 (68 % of total OC) to 1.4 ± 0.1 µg m−3 (60 % of total OC) between 1999 and 2013 at the urban Atlanta, Georgia, site (JST) and from 2.9 ± 0.4 µg m−3 (39 % of total OC) to 0.7 ± 0.1 µg m−3 (30 % of total OC) between 2001 and 2013 at the urban Birmingham, Alabama (BHM), site. The urban OC declines coincide with reductions of motor vehicle emissions between 2006 and 2010, which may have decreased mean OC concentrations at the urban SEARCH sites by > 2 µg m−3. BHM additionally exhibits a decline in OC associated with SO2 from 0.4 ± 0.04 µg m−3 in 2001 to 0.2 ± 0.03 µg m−3 in 2013, interpreted as the result of reduced emissions from industrial sources within the city. Analyses using non-soil potassium as a biomass burning tracer indicate that biomass burning OC occurs throughout the year at all sites. All eight SEARCH sites show an association of OC with sulfate (SO4) ranging from 0.3 to 1.0 µg m−3 on average, representing ∼ 25 % of the 1999 to 2013 mean OC concentrations. Because the mass of OC identified with SO4 averages 20 to 30 % of the SO4 concentrations, the mean SO4-associated OC declined by ∼ 0.5 to 1 µg m−3 as SO4 concentrations decreased throughout the SEARCH region. The 2013 mean SO4 concentrations of 1.7 to 2.0 µg m−3 imply that future decreases in mean SO4-associated OC concentrations would not exceed ∼ 0.3 to 0.5 µg m−3. Seasonal OC concentrations, largely identified with ozone (O3), vary from 0.3 to 1.4 µg m−3 ( ∼ 20 % of the total OC concentrations).
Article
Full-text available
Long-term (1999 to 2013) data from the Southeastern Aerosol Research and Characterization (SEARCH) network are used to characterize the effects of anthropogenic emission reductions on fine particle organic aerosol (OA) concentrations in the southeastern US. On average, 45 % (range 25 to 63 %) of the 1999 to 2013 mean organic carbon (OC) concentrations are attributed to combustion processes, including fossil-fuel use and biomass burning, through associations of measured OC with combustion products such as elemental carbon (EC), carbon monoxide (CO), and nitrogen oxides (NOx). The 2013 mean combustion-derived OC concentrations were 0.5 to 1.4 μg m−3 at the five sites operating in that year. Mean annual combustion-derived OC concentrations declined from 3.8 ± 0.2 μg m−3 (68 % of total OC) to 1.4 ± 0.1 μg m−3 (60 % of total OC) between 1999 and 2013 at the urban Atlanta, Georgia, site (JST) and from 2.9 ± 0.4 μg m−3 (39 % of total OC) to 0.7 ± 0.1 μg m−3 (30 % of total OC) between 2001 and 2013 at the urban Birmingham, Alabama, site (BHM). The urban OC declines coincide with reductions of motor-vehicle emissions between 2006 and 2010, which may have decreased mean OC concentrations at the urban SEARCH sites by > 2 μg m−3. BHM additionally exhibits a decline in OC associated with SO2 from 0.4 ± 0.04 μg m−3 in 2001 to 0.2 ± 0.03 μg m−3 in 2013, interpreted as the result of reduced emissions from industrial sources within the city. Analyses using non-soil potassium as a biomass-burning tracer indicate that biomass-burning OC occurs throughout the year at all sites. All eight SEARCH sites show an association of OC with sulfate (SO4) ranging from 0.3 to 1.0 μg m−3 on average, representing ~ 25 % of the 1999 to 2013 mean OC concentrations. Because the mass of OC associated with SO4 averages 20 to 30 % of the SO4 concentrations, the mean SO4-associated OC declined by ~ 0.5 to 1 μg m−3 as SO4 decreased throughout the SEARCH region. The 2013 mean SO4 concentrations of 1.7 to 2.0 μg m−3 imply that future decreases in mean SO4-associated OC concentrations would not exceed ~ 0.3 to 0.5 μg m−3. Seasonal OC concentrations, largely associated with ozone (O3), vary from 0.3 to 1.4 μg m−3 (~ 20 % of the total OC concentrations).
Article
Full-text available
Methacryloyl peroxynitrate (MPAN), the acyl peroxynitrate of methacrolein, has been suggested to be an important secondary organic aerosol (SOA) precursor from isoprene oxidation. Yet, the mechanism by which MPAN produces SOA via reaction with the hydroxyl radical (OH) is unclear. We systematically evaluate three proposed mechanisms in controlled chamber experiments and provide the first experimental support for the theoretically-predicted lactone formation pathway from the MPAN + OH reaction, producing hydroxymethyl-methyl-α-lactone (HMML). The decomposition of the MPAN-OH adduct yields HMML + NO3 (~ 75%) and hydroxyacetone + CO + NO3 (~ 25%), out-competing its reaction with atmospheric oxygen. The production of other proposed SOA precursors, e.g., methacrylic acid epoxide (MAE), from MPAN and methacrolein are negligible (< 2 %). Furthermore, we show that the beta-alkenyl moiety of MPAN is critical for lactone formation. Alkyl radicals formed cold via H-abstraction by OH not decompose to HMML, even if they are structurally identical to the MPAN-OH adduct. The SOA formation from HMML, via polyaddition of the lactone to organic compounds at the particle interface or in the condensed phase, is close to unity under dry conditions. However, the SOA yield is sensitive to particle liquid water and solvated ions. In hydrated inorganic particles, HMML reacts primarily with H2O to produce the monomeric 2-methylglyceric acid (2MGA) or aqueous sulfate and nitrate to produce the associated organosulfate and organonitrate, respectively. 2MGA, a tracer for isoprene SOA, is semivolatile and its accommodation in aerosol water decreases with decreasing pH. Conditions that enhance the production of neutral 2MGAsuppress SOA mass from the HMML channel. Considering the liquid water content and pH ranges of ambient particles, 2MGA will exist largely as a gaseous compound in some parts of the atmosphere.
Article
Full-text available
A suite of offline and real-time gas- and particle-phase measurements was deployed at Look Rock, Tennessee (TN), during the 2013 Southern Oxidant and Aerosol Study (SOAS) to examine the effects of anthropogenic emissions on isoprene-derived secondary organic aerosol (SOA) formation. High- and low-time resolution PM2.5 samples were collected for analysis of known tracer compounds in isoprene-derived SOA by gas chromatography/electron ionization-mass spectrometry (GC/EI-MS) and ultra performance liquid chromatography/diode array detection-electrospray ionization-high-resolution quadrupole time-of-flight mass spectrometry (UPLC/DAD-ESI-HR-QTOFMS). Source apportionment of the organic aerosol (OA) was determined by positive matrix factorization (PMF) analysis of mass spectrometric data acquired on an Aerodyne Aerosol Chemical Speciation Monitor (ACSM). Campaign average mass concentrations of the sum of quantified isoprene-derived SOA tracers contributed to ~9% (up to 26%) of the total OA mass, with isoprene-epoxydiol (IEPOX) chemistry accounting for ~97% of the quantified tracers. PMF analysis resolved a factor with a profile similar to the IEPOX-OA factor resolved in an Atlanta study and was therefore designated IEPOX-OA. This factor was strongly correlated (r2>0.7) with 2-methyltetrols, C5-alkene triols, IEPOX-derived organosulfates, and dimers of organosulfates, confirming the role of IEPOX chemistry as the source. On average, IEPOX-derived SOA tracer mass was ~25% (up to 47%) of the IEPOX-OA factor mass, which accounted for 32% of the total OA. A low-volatility oxygenated organic aerosol (LV-OOA) and an oxidized factor with a profile similar to 91Fac observed in areas where emissions are biogenic-dominated were also resolved by PMF analysis, whereas no primary organic aerosol (POA) sources could be resolved. These findings were consistent with low levels of primary pollutants, such as nitric oxide (NO~0.03ppb), carbon monoxide (CO~116 ppb), and black carbon (BC~0.2 μg m−3). Particle-phase sulfate is fairly correlated (r2~0.3) with both MAE- and IEPOX-derived SOA tracers, and more strongly correlated (r2~0.6) with the IEPOX-OA factor, in sum suggesting an important role of sulfate in isoprene SOA formation. Moderate correlation between the methacrylic acid epoxide (MAE)-derived SOA tracer 2-methylglyceric acid with sum of reactive and reservoir nitrogen oxides (NOy; r2=0.38) and nitrate (r2=0.45) indicates the potential influence of anthropogenic emissions through long-range transport. Despite the lack of a~clear association of IEPOX-OA with locally estimated aerosol acidity and liquid water content (LWC), box model calculations of IEPOX uptake using the simpleGAMMA model, accounting for the role of acidity and aerosol water, predicted the abundance of the IEPOX-derived SOA tracers 2-methyltetrols and the corresponding sulfates with good accuracy (r2~0.5 and ~0.7, respectively). The modeling and data combined suggest an anthropogenic influence on isoprene-derived SOA formation through acid-catalyzed heterogeneous chemistry of IEPOX in the southeastern US. However, it appears that this process was not limited by aerosol acidity or LWC at Look Rock during SOAS. Future studies should further explore the extent to which acidity and LWC becomes a limiting factor of IEPOX-derived SOA, and their modulation by anthropogenic emissions.
Article
Isoprene is the most abundant non-methane hydrocarbon emitted into Earth’s atmosphere and is predominantly derived from terrestrial vegetation. Prior studies have focused largely on the hydroxyl (OH) radical-initiated oxidation of isoprene and have demonstrated that highly oxidized compounds, such as isoprene-derived epoxides, enhance the formation of secondary organic aerosol (SOA) through heterogeneous (multiphase) reactions on acidified sulfate aerosol. However, studies on the impact of acidified sulfate aerosol on SOA formation from isoprene ozonolysis are lacking and the current work systematically examines this reaction. SOA was generated in an indoor smog chamber from isoprene ozonolysis under dark conditions in the presence of non-acidified or acidified sulfate seed aerosol. The effect of OH radicals on SOA chemical composition was investigated using diethyl ether as an OH radical scavenger. Aerosols were collected and chemically characterized by ultra performance liquid chromatography/electrospray ionization high-resolution quadrupole time-of-flight mass spectrometry (UPLC/ESI-HR-QTOFMS) and gas chromatography/electron impact ionization-mass spectrometry (GC/EI-MS). Analysis revealed the formation of highly oxidized compounds, including organosulfates (OSs) and 2-methylterols, which were significantly enhanced in the presence of acidified sulfate seed aerosol. OSs identified in the chamber experiments were also observed and quantified in summertime fine aerosol collected from two rural locations in the southeastern United States during the 2013 Southern Oxidant and Aerosol Study (SOAS).