Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Atmospheric Environment 225 (2020) 117375

Contents lists available at ScienceDirect

Atmospheric Environment
journal homepage: http://www.elsevier.com/locate/atmosenv

NOx enhances secondary organic aerosol formation from nighttime


γ-terpinene ozonolysis
Li Xu a, Narcisse T. Tsona b, Bo You a, Yingnan Zhang a, Shuyan Wang a, Zhaomin Yang a,
Likun Xue a, Lin Du a, *
a
Environment Research Institute, Shandong University, Qingdao, 266237, China
b
School of Life Science, Shandong University, Qingdao, 266237, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

� Ozonolysis of γ-terpinene is not an


important SOA source in NOx-isolated
regions.
� SOA size and yields from γ-terpinene
ozonolysis increase with the increasing
NOx.
� Organic nitrates are important compo­
nents of SOA from NOx-involved
ozonolysis.
� Nitrate radical contributes to nighttime
γ-terpinene oxidation.

A R T I C L E I N F O A B S T R A C T

Keywords: As an important component of anthropogenic emissions, nitrogen oxides (NOx) are well known to interfere with
γ-Terpinene daytime oxidation of biogenic volatile organic compounds (BVOCs) and secondary organic aerosol (SOA) for­
Ozonolysis mation. Nighttime chemistry is highly related to O3 and NOx level but, NOx effects on SOA formation from the
Secondary organic aerosol
ozonolysis of BVOCs, especially polyolefinic monoterpenes, have not been well understood. In the present study,
Nighttime
Anthropogenic-biogenic interactions
SOA formation from pure and NOx-involved γ-terpinene ozonolysis was studied in a smog chamber under dark
conditions. At atmospherically relevant particle mass loading of 10 μg m 3, the SOA yield from pure ozonolysis is
estimated by a two-product model to be 8.6%. When NOx were incorporated into γ-terpinene ozonolysis, both
the particle size and SOA yields increased simultaneously with elevated NOx mixing ratios. SOA yields doubled
(from 0.38 to 0.77) when the system moved from NOx-free to [γ-terpinene]0/[NOx]0 ¼ 3.5 (ppbC/ppb). The
characteristic absorption of organic nitrates was detected by Fourier transform infrared (FTIR) spectroscopy and
the fraction of organic nitrates increased with increasing NOx mixing ratios. Identification of the new constit­
uents in SOA from NOx-involved γ-terpinene ozonolysis and their formation channels suggest that the formation
of organic nitrates follows NO3 chemistry. NOx affects γ-terpinene ozonolysis via the enhanced generation of NO3
at high NOx and its subsequent more favored consumption of γ-terpinene than O3. The first-generation products
from NO3 oxidation of γ-terpinene could be further oxidized by ozone, forming more oxidized products that
contribute to SOA formation. Our investigation suggests that at night with high NOx levels, γ-terpinene may be a
significant source of SOA and organic nitrates through anthropogenic-biogenic interactions.

* Corresponding author.
E-mail address: lindu@sdu.edu.cn (L. Du).

https://doi.org/10.1016/j.atmosenv.2020.117375
Received 18 October 2019; Received in revised form 21 February 2020; Accepted 24 February 2020
Available online 27 February 2020
1352-2310/© 2020 Elsevier Ltd. All rights reserved.
L. Xu et al. Atmospheric Environment 225 (2020) 117375

1. Introduction exocyclic double bond) (Zhao et al., 2015). To the best of our knowl­
edge, NOx effects on dark ozonolysis of other monoterpenes (e.g. with
The oxidation of volatile organic compounds (VOCs) generates two endocyclic double bonds) have been rarely studied.
products that undergo chemical transformation and condensation to γ–Terpinene is one of monoterpenes with two substituted endocyclic
form secondary organic aerosols (SOA) (Jimenez et al., 2009; Zhang double bonds. Measurements in a tropical cloudy forest in Costa Rica
et al., 2007). SOA contribute to a major fraction of submicron atmo­ showed that γ–terpinene mixing ratio is up to 1.5 ppb in October, higher
spheric aerosols, which are highly linked to the visibility reduction, than those of β-pinene (0.4 ppb) and limonene (1.3 ppb) (Esqui­
Earth’s radiation balance and regional climate change as well as human vel-Herna �ndez et al., 2011). In a rural mixed forest in northern Michi­
respiratory and cardiovascular diseases (Hallquist et al., 2009; Jimenez gan, the simulated formation of organic nitrates from the reaction of
et al., 2009; Kavouras et al., 1998; Mauderly and Chow, 2008; Zaveri γ–terpinene with NO3 was comparable to those of limonene, α-pinene,
et al., 2018). Biogenic VOCs (BVOCs) emitted by vegetation comprises and β-pinene (Pratt et al., 2012). While SOA formation from α-pinene,
90% of the global VOC emissions (Guenther et al., 1995). Biogenic SOA β-pinene and limonene oxidation have been widely studied under
(BSOA) derived from BVOCs oxidation are estimated to be more than various conditions, γ–terpinene oxidation has received less attention
one order of magnitude larger than SOA formed from anthropogenic (Friedman and Farmer, 2018; Griffin et al., 1999; Slade et al., 2017).
VOCs (Sakulyanontvittaya et al., 2008; Virtanen et al., 2010). The SOA yields from γ–terpinene photooxidation were measured to be
importance of anthropogenic emissions in SOA formation is reflected in 0.10–0.16 within the aerosol mass range of 21.3–65.9 μg cm 3 (Griffin
their modulation of BVOCs oxidation and consequent BSOA formation as et al., 1999). Using N2O5 as NO3 radicals precursor, γ–terpinene
many field observations and modeling predictions suggested (Budisu­ oxidation produced a total organic nitrates yield of 14% and SOA yield
listiorini et al., 2015; Han et al., 2016; Liu et al., 2018; Shrivastava et al., up to 10% under an atmospherically relevant aerosol mass loading of 10
2017; Spracklen et al., 2011). μg cm 3 (Slade et al., 2017). These studies suggest that γ–terpinene can
Nitrogen oxides (NOx) are important anthropogenic pollutants that be an important BSOA contributor, especially in regions with high
are mainly emitted from combustion-related activities (Draper et al., γ–terpinene emission rates, such as Midwest USA where γ–terpinene
2015; Reis et al., 2009). They have been receiving significant attention makes up ~5% of the monoterpene emissions (Geron et al., 2000).
due to their effects on the chemical evolution of BVOCs and SOA for­ In this study, we investigated SOA formation from γ–terpinene ozo­
mation (Berkemeier et al., 2016; Ng et al., 2017; Renbaum and Smith, nolysis and the yield data were fitted by a two-product model. NOx ef­
2009; Wildt et al., 2014). Recently, while many model studies showed fects on SOA yield and size distribution from γ–terpinene ozonolysis
the increase of BSOA with decreased NOx, some studies demonstrated were first examined in the present study. The chemical composition of
that NOx reduction may result in the decrease of BSOA formation (Lane SOA formed under different NOx conditions was also explored with
et al., 2008; Matsui et al., 2014; Pye et al., 2013; Zhang et al., 2017). For Fourier transform infrared spectrum and high-resolution mass spec­
example, based on semi-volatile partitioning mechanism for SOA for­ trometry. Considering that the nighttime atmospheric oxidation capac­
mation modeled by Community Multiscale Air Quality model, BSOA ity is highly related to O3 and NOx level, we used the NOx/O3 ratio of
increased by 5% when NOx was reduced by 25% in Southeastern 0–0.7 that is relevant to the nighttime ambient conditions to assess the
America (Pye et al., 2013). A NOx reduction of 50% resulted in a limited SOA formation potential of γ–terpinene through anthropogenic-biogenic
SOA reduction of 0.9–5.6% on a global scale as modeled by Community interactions at night.
Atmospheric Model version 4 with chemistry (Zheng et al., 2015). These
multifaceted effects of NOx on BSOA formation in the atmosphere are 2. Experimental methods
related to the specific atmospheric conditions and parent precursors (Liu
et al., 2018). 2.1. SOA generation
BVOC oxidation induced by nitrate radical (NO3) is deemed as one of
the anthropogenic-biogenic interactions because nighttime NO3 radicals All experiments were carried out in a collapsible 1 m3 Teflon
formation occurs predominantly via the reaction of NOx with ozone chamber under dark conditions and ambient pressure. The initial
(Brown and Stutz, 2012; Kiendler-Scharr et al., 2016). In the USA, it was experimental conditions are summarized in Table 1. Based on the
estimated that monoterpene oxidation by the NO3 radical may lead to different experimental purposes, these experiments can be grouped into
more than half of the SOA derived from monoterpene oxidation (Pye three categories: pure ozonolysis of γ–terpinene (Exp.1-8), NOx-
et al., 2015). Other studies showed that the nighttime contribution of involved ozonolysis with initial ~153 ppb of γ–terpinene
organic nitrates (4%–8%) to the total organic aerosol particle was higher ([γ–terpinene]0 ¼ ~153 ppb, Exp.9-15) and initial ~302 ppb of
than that during the daytime (2%–4%) (Lee et al., 2016; Rollins et al., γ–terpinene ([γ–terpinene]0 ¼ 302 ppb, Exp.16-26). The temperature
2012). Laboratory experiments usually use N2O5 as the source of NO3 and relative humidity (RH) inside the chamber were monitored with a
radicals to study SOA and organic nitrates formation from NO3-initiated hygrometer (Model 645, Testo AG, Germany). Before each experiment,
BVOCs oxidation (Boyd et al., 2015; Fry et al., 2014; Griffin et al., 1999; the chamber was flushed continuously with zero air that was generated
Moldanova and Ljungstro €m, 2000; Perraud et al., 2010). In fewer studies by a zero air supply (111-D3N, Thermo Scientific, USA) until the particle
NOx and ozone were introduced into the chamber to study NOx effects number concentrations were less than 10 cm 3.
on the dark ozonolysis of BVOCs. NOx effects on SOA formation from the The introduction order of reagents was γ–terpinene, cyclohexane,
ozonolysis of several commonly studied monoterpenes have been found NOx (in NOx-involved experiments) and ozone. A known volume of
to be different (Draper et al., 2015; Nojgaard et al., 2006; Presto et al., liquid γ–terpinene (Aldrich, 97%) was first injected into a stream of zero
2005b). In α-pinene ozonolysis under dark conditions, characterized by air and evaporated into the chamber directly. Then, an excess amount of
[VOC]0/[NOx]0 ratio varying from ~1 to ~300, the SOA yield from 150 cyclohexane (Aladdin, 99.5%) was flushed into the chamber by zero air
to 200 ppb α-pinene ozonolysis remained constant (~0.30) for so that more than 90% of the OH radical formed by ozone reaction with
[VOC]0/[NOx]0 > 15 and it decreased to be lower than 0.01 for γ–terpinene reaction could be scavenged (Lee et al., 2006; Nah et al.,
[VOC]0/[NOx]0 ¼ 0.65 (Presto et al., 2005b). With [O3]/[NO2] ranging 2016). In experiments where NOx were involved, an expected amount of
from 1:0.5 to 1:4 and [NO2]/[VOC] ranging from 1:1 to 2:1, β-pinene NO or NO2 from the calibrated cylinder (490 ppm, YuyanGas®, China)
and Δ3-carene produced fewer particles while limonene yielded higher was introduced into the chamber with a gas-tight syringe. These gas
aerosol number and mass concentrations at high NOx levels (Draper components were allowed to stabilize for 20 min before the injection of
et al., 2015). The varied SOA formation behavior of these monoterpenes ozone that was produced by an ozone generator (WH-H-Y5Y, Wohuan,
may be related to their different isomeric structures (e.g., α-pinene China). The reaction in the chamber was initiated by ozone and pro­
contains one endocyclic double bond and β-pinene contains one ceeded for two hours before the collection of SOA particles. Control

2
L. Xu et al. Atmospheric Environment 225 (2020) 117375

experiments showed negligible effects of cyclohexane on NOx chemistry in the first two hours. Although neglecting organic vapor wall loss may
and SOA formation. cause an underestimation of SOA yields, the wall loss of organic vapor in
A set of real-time instruments were employed to characterize the gas- the present study may be mitigated by the use of excess oxidant con­
phase and particle-phase components. The concentrations of NOx and centrations. The quick ozonolysis and nucleation products would pro­
ozone were continuously measured by the NO–NO2-NOx analyzer vide surfaces for vapors to partition onto, therefore reducing the vapor
(Model 42i, Thermo Scientific, USA), and O3 analyzer (Model 49i, wall loss (Nah et al., 2016; Stirnweis et al., 2017). Particle number and
Thermo Scientific, USA), respectively. Their sampling flow rate was 0.7 mass concentrations in the first two hours were corrected for wall loss
L min 1. No air was added into the chamber to avoid dilution and to based on the size-dependent particle loss rates in each experiment (Chen
maintain a relatively stable chamber condition during the reaction. The et al., 2019a; Chu et al., 2016; Takekawa et al., 2003). Details about this
γ-terpenene concentration that was detected at 30 min interval by a gas method are described in Fig. S1 in the supplementary material. Wall
chromatography-flame ionization detector (GC-FID, 7890B, Agilent effect in the particle collection period has not been evaluated as it would
Technologies, USA) with a DB-624 capillary column (30 m � 0.32 mm, have minor effect on the SOA composition analysis.
1.8 mm film thickness) could not be detected when SOA sampling
started. The particle number concentration and size distribution were
2.2. SOA collection and composition analysis
monitored throughout each experiment using the scanning mobility
particle sizer (SMPS, model 3080, TSI Inc., USA). SMPS was operated
Sampling started after two hours after the reaction was initiated. The
with a sheath flow of 3 L min 1 and aerosol flow of 0.3 L min 1. Particle
remaining ~0.8 m3 mixture in the chamber was enough for approxi­
sizes from 13.6 to 736.5 nm were recorded with a scanning cycle of 5
mately 1 h sampling at 10 L min 1 flow rate, during which some re­
min. Since these experiments were performed in the absence of seed
actions might still proceed. SOA particles generated in experiments with
particles that can provide surfaces for vapors to condense onto, the
initial ~302 ppb γ–terpinene (Exp. 16-24) were collected onto
initial particles formation in the chamber was driven by nucleation of
aluminum foil pieces by a 13-stage Dekati low-pressure impactor
oxidation products.
(DLPIþ, DeKati Ltd, Finland). Functional groups and bond information
The mass concentration of SOA generated in each reaction system
of particle-phase products were analyzed using attenuated total reflec­
was also detected by SMPS with an estimated aerosol density of 1.2 g
tance coupled to the Fourier transform infrared spectroscopy (ATR-
cm 3. The mass-dependent SOA yield, representative of SOA formation
FTIR, Vertex 70, Bruker, Germany). The sample spectra in the range of
potential of a parent hydrocarbon, is defined as the consumption of
4000-600 cm 1 were scanned by placing the aluminum foil pieces
γ-terpinene (Δ[γ-terpinene], μg m 3) that is converted to SOA mass
directly onto the ATR crystal with the resolution of 4 cm 1 and 64 scans
(ΔM, μg m 3). It should be noted that both vapor and particles could be
were averaged. To ensure a clean background, the ATR crystal was
lost to the chamber wall and influence SOA yields (Krechmer et al.,
thoroughly cleaned after each sample spectrum acquisition, and the
2016; Ye et al., 2016; Zhang et al., 2015). The wall loss rate constants for
aluminum foil spectrum as the background was subtracted. The spectra
NOx, ozone and γ-terpinene were measured to be 1.72 � 10 7, 2.24 �
baseline was corrected using the rubberband method (20 iterations and
10 6, 6.8 � 10 7 s 1, respectively. These wall losses were less than 3%
64 baseline points) in OPUS software.

Table 1
Experimental conditions.
Exp. [γ-terpinene]0 [O3]0 [NOx]0 [γ-terpinene]0/[NOx]0 Temperature RH SOA
Yield
No. (ppb) (ppb) (ppb) (ppbC/ppb) (K) (%)

1a 31 123 0 – 297 34 0.14


2a 60 216 0 – 297 30 0.25
3a 76 304 0 – 300 27 0.29
4a 122 488 0 – 298 29 0.35
5a 154 629 0 – 299 30 0.38
6a 183 709 0 – 297 30 0.48
7a 215 763 0 – 298 31 0.55
8a 276 983 0 – 298 30 0.56

9b 154 565 37 40.6 300 29 0.40


10b 153 557 55 27.8 298 30 0.49
11b 155 602 76 20.3 301 25 0.52
12b 153 578 86 17.8 298 29 0.55
13b 152 609 144 10.7 297 36 0.65
14b 154 610 310 5.0 299 24 0.67
15b 153 590 441 3.5 299 30 0.77

16c 302 1236 0 – 292 17 0.55


17c 302 942 28 107.9 297 21 0.61
18c 302 1141 59 51.4 294 19 0.72
19c 302 992 145 20.8 293 20 0.82
20c 302 1250 238 12.7 294 19 0.92
21c 302 1430 461 6.6 294 20 0.97
22c 302 1475 550 5.5 295 22 1.09
23c 302 1611 603 5.0 294 19 1.18
24c 302 1780 780 3.9 294 23 1.30

25d 302 937 0 – 295 21 –


26d 302 935 547 5.5 293 22 –
a
Experiments aiming at studying SOA yields from pure ozonolysis.
b
Experiments aiming at studying SOA size and yields in NOx-involved ozonolysis: NO2 was added.
c
SOA size and yield in NOx-involved ozonolysis and particles were collected onto aluminum foil for composition analysis by FTIR: NO was added. The addition of O3
was adjusted with the mixing ratio of NO to compensate the quick consumption of O3 by NO.
d
Particles were collected onto PTFE filters for composition analysis by UPLC/ESI-HR-Q-TOFMS.

3
L. Xu et al. Atmospheric Environment 225 (2020) 117375

SOA composition characterization was also performed using an


ultra-performance liquid chromatography (UPLC, Ultimate 3000,
Thermo scientific, USA) system coupled to the electrospray ionization
high-resolution quadrupole time-of-flight mass spectrometry (ESI-HR-Q-
TOFMS, Bruker Impact HD, Germany). SOA (Exp. 25-26) were collected
onto polytetrafluoroethylene (PTFE) filters (0.22 μm pore size, 47 mm
dimeter, TJMF50, Jinteng, China) by a stainless steel 47 mm inline filter
holder (Sartorius 16254, Sartorius Stedim Biotech GmbH, Germany).
PTFE samples were first extracted with 5 mL methanol (Optima® LC/MS
grade, Fisher Scientific) by sonication for 30 min. Then the extracts were
filtered through a 25 mm diameter, 0.22 μm pore PTFE membrane sy­
ringe filter (Jinteng, China) to remove insoluble material. After the
extracted solution was concentrated to almost dryness under a gentle
stream of nitrogen, the extract residues were reconstituted with 500 μL
of 50:50 (v/v) solvent mixture of 0.1% formic acid (Optima® LC/MS
grade, Fisher Scientific) in methanol and 0.1% formic acid in ultrapure
water (Milli-Q, Millipore, France). The blank controls of filters were
handled and analyzed as samples to avoid the interference of PTFE filter
artifacts on composition analysis. Finally, these prepared samples were
analyzed by UPLC/ESI-HR-Q-TOFMS that was operated in negative
ionization mode with a mass range of m/z 50-1000. Chromatographic
separations were performed using an Atlantis T3 C18 column (100 Å, 3
μm particle size, 2.1 mm � 150 mm, Waters, USA). The mobile phases
consisted of (A) 0.1% formic acid in water and (B) 0.1% formic acid in Fig. 1. Time evolution of wall-loss corrected SOA size (aerodynamic diameter)
methanol. The applied 60 min gradient elution program, with the flow and mass concentration obtained from γ-terpinene/O3 experiment (Exp. 6).
rate of 200 μL min 1 and injection volume of 5 μL, was as follows: 3% B
for the first 3 min, increased linearly to 50% B from 3 to 25 min, and
increased from 50% to 90% B for another 18 min, decreased back to 3%
B from 43 to 48 min, and then held at 3% B from 48 to 60 min
(Kourtchev et al., 2015).

3. Results and discussion

3.1. γ-Terpinene ozonolysis

SOA formation from γ–terpinene ozonolysis was firstly investigated


at different initial γ–terpinene concentrations. The time evolution pro­
file of particle wall loss corrected SOA number concentration and total
mass concentration are plotted in Fig. 1 (Exp. 6). Particle formation is
observed by the burst of the particle number in the first few minutes of
ozone injection. These fast-formed small particles grow quickly to be
larger than 100 nm within 20 min with a decrease in number concen­
tration, which may appear if the coagulation between these small par­
ticles and/or the partitioning of gas-phase products to particle-phase
exceed the new particle formation through homogeneously nucleation
(Friedman and Farmer, 2018). The SOA mass concentration reaches its
maximum at about one hour after the initiation of the reaction. In the Fig. 2. SOA yields of γ-terpinene ozonolysis as a function of SOA mass con­
absence of seeds, the supersaturation of condensable products that have centrations (ΔM) (Exp.1-8). The experimental data are fitted with the two-
extremely low volatility was suggested to drive the initial nucleation product model. The mass-dependent yield data of α-pinene, β-pinene, and
(Donahue et al., 2013). The formation of extremely low volatility limonene ozonolysis reported in previous studies were normalized to density ¼
organic compounds (ELVOCs) was observed within the first minute after 1.2 g cm 3 for comparison. The experimental conditions for these ozonolysis
the injection of ozone to monoterpenes ozonolysis systems (Jokinen experiments are summarized in Table S1.
et al., 2015). Endocyclic BVOCs were thus concluded to be efficient
precursors for ELVOCs formation through ozonolysis. The burst of X αi Kom;i
nanoparticles in γ-terpinene/O3 system at the initial few minutes may Y ¼ M0
1 þ Kom;i M0
also result from the quick formation of ELVOCs and their homogeneous i

nucleation. While the formation and nucleation of ELVOCs may explain


where αi is the mass-based stoichiometric fraction of product i and Kom,i
the primary increase of SOA mass loadings, semi-volatile organic com­
(m3 μg 1) is the gas–particle partitioning equilibrium constant. In this
pounds (SVOCs) probably begin to contribute more through absorptive
model, one group of parameters (α1 and Kom,1) represents the ensemble
partitioning at higher SOA mass loadings (~10 μg m 3) (Ehn et al.,
of products with relatively low volatility and the other group of pa­
2014).
rameters (α2 and Kom,2) designates the ensemble of products with rela­
SOA yields increase with the increasing total SOA mass loadings as
tively high volatility. The experimental values of SOA yields versus mass
shown in Fig. 2. This correlation can be expressed by a semi-empirical
concentrations can be well reproduced by the two-product model as the
model, which is built on the basis of the absorptive gas-particle parti­
well-fitted curve (R2 ¼ 0.99) shown in Fig. 2. The final fitting parameters
tioning of produced SVOCs (Odum et al., 1996).
are α1 ¼ 0.77, Kom,1 ¼ 7.64 � 10 4, α2 ¼ 0.23, and Kom,2 ¼ 0.05. Taking

4
L. Xu et al. Atmospheric Environment 225 (2020) 117375

a typical 10 μg m 3 of aerosol mass loading in regions affected by BSOA


into consideration, the simulated SOA yield is 8%. At such a lower
aerosol mass loading, the rapid SOA yield increase is most likely caused
by the partitioning of the lowest volatility products. When the aerosol
mass loading is larger than 500 μg m 3, which is relevant to heavily
polluted environments, the SOA yield reaches 0.47. The SOA yield
change became slower at high SOA mass concentrations, suggesting the
partitioning of relative volatile species.
The ozonolysis of some other monoterpenes has been well studied
previously (Griffin et al., 1999; Lee et al., 2006; Saathoff et al., 2009;
Song et al., 2007; von Hessberg et al., 2009). For comparison of SOA
formation potential of different isomeric monoterpenes, the results from
α-pinene, β-pinene, and limonene ozonolysis conducted under compa­
rable conditions are also plotted in Fig. 2, while their experimental
conditions are summarized in Table S1. As shown in Fig. 2, the reaction
between limonene and ozone (293 K, dry, Saathoff et al.) has the highest
yields, including when temperature and RH effects are considered (the
~5 K higher temperature may cause the SOA yield to decrease by 5%
and humidity was suggested to have little effect on SOA yields from Fig. 3. Size distributions of SOA particles generated under different NOx levels
limonene ozonolysis) (Presto et al., 2005a; Saathoff et al., 2009; von at 1 h after the initiation of the reaction.
Hessberg et al., 2009). SOA yields from α-pinene ozonolysis that was
performed with RH < 2% by Song et al. are also higher than that in and 11 nm for d-limonene compared to pure ozonolysis, and the particle
γ-terpinene ozonolysis under RH ¼ ~30% (Song et al., 2007). The higher number concentration decreased in both α-pinene and d-limonene cases
α-pinene SOA yield is still expected even under similar humidity con­ (Nojgaard et al., 2006). It was explained that introducing NOx into the
dition because humidity would not influence SOA formation from reaction system resulted in less O3 available for the oxidation of parent
α-pinene ozonolysis (Saathoff et al., 2009). β-Pinene ozonolysis that was compounds and, thus, to the decreased particle number concentration.
conducted under dry conditions by Hessberg et al. exhibited lowest SOA However, this might not be the case for γ-terpinene ozonolysis because
yields compared to limonene, α-pinene, and γ-terpinene (von Hessberg ozone is in excess in our experiments. The presence of NOx tends to
et al., 2009). Considering that RH would suppress β-pinene SOA for­ affect the gas-phase reaction mechanism in different ways among which,
mation, SOA yields from the γ-terpinene þ O3 reaction are expected to one is to react with organic peroxy radicals (RO2). The RO2 þ NO re­
be comparable to β-pinene ozonolysis in low aerosol mass loading and action is insignificant in the present study because the pre-existing NO
much larger when the aerosol mass loading is greater than 100 μg m 3 can be rapidly converted to NO2 once excessive ozone is injected into the
(Moortgat, 2002). chamber. RO2 reacts with NO2 to produce peroxy nitrates (RO2NO2),
Different SOA formation potential from isomeric monoterpenes can which are thermally labile under ambient conditions and less likely to
be attributed to their different structures (Griffin et al., 1999; Zhao et al., contribute to SOA formation (Atkinson, 2007; Berkemeier et al., 2016;
2015). It has been suggested that the ozonolysis of monoterpenes with Hallquist et al., 1999). Another fate of NOx in our experiments is to form
endocyclic double bonds, such as α-pinene, produced ELVOCs very NO3 radicals by the reaction between O3 and NO2. Monoterpenes
rapidly with much greater molar yields than monoterpenes with oxidation by NO3 radical is potentially efficient for SOA and organic
exocyclic double bonds, such as β-pinene (Jokinen et al., 2015). Limo­ nitrates formation (Fry et al., 2014; Ng et al., 2017). If products from
nene has one endocyclic and one exocyclic double bond in its structure. quick oxidation of γ-terpinene by NO3 radical and/or following sec­
The unsaturated first-generation products can further react with O3 to ondary reactions (e.g. the further ozonolysis of the first-generation
form highly oxidized products with high molecular weight that can products) are much easier to partition into the particle-phase, the par­
easily condense onto the particle phase. Hence, the SOA yields from ticle formation would be promoted. The following particle growth
limonene ozonolysis are the highest (Saathoff et al., 2009). Despite the through coagulation can result in the increase of particle size and
two endocyclic double bonds in γ-terpinene, the SOA yields from its decrease of particle number concentration.
ozonolysis are lower than in that of limonene. Because the ozonolysis of SOA yields from γ-terpinene ozonolysis at varying [γ-terpinene]0/
the second double bond in γ-terpinene may cause the dissociation of the [NOx]0 ratios are plotted in Fig. 4. They vary quite dramatically as the
carbon skeleton, the consequent products with smaller molecular weight reaction system moves from low NOx to high NOx mixing ratios.
might be more volatile and thus the SOA yield would be lower (Lee et al., Compared to pure ozonolysis where the SOA yield is 0.38, the SOA yield
2006). increases to 0.77 when NOx level is elevated to [γ-terpinene]0/[NOx]0
¼ 3.5 in experiments with initial ~ 153 ppb γ-terpinene. Even for similar
3.2. Effect of NOx on γ-terpinene ozonolysis [γ-terpinene]0/[NOx]0 ratios, significantly higher SOA yields are
observed in experiments with initial higher γ-terpinene mixing ratios
3.2.1. Effect of NOx on SOA size distribution and yields (~302 ppb). Positive effects of precursor concentrations on SOA yields
The size distribution of particles from different [γ-terpinene]0/ were found in NOx associated α-pinene ozonolysis. With a [VOC]0/
[NOx]0 (ppbC/ppb) systems at 1 h after the initiation of the reaction is [NO2]0 ratio of ~15, the SOA yields of 0.3 and only 0.08 were observed
shown in Fig. 3. In two cases with different initial γ-terpinene mixing for experiments with incipiently higher (150–200 ppb) and lower
ratios, the decrease of [γ-terpinene]0/[NOx]0 ratios reduced particle (15–30 ppb) α-pinene mixing ratios, respectively (Presto et al., 2005b).
number concentrations compared with pure ozonolysis, concurrent with However, NOx exhibit quite different effects on SOA yields from
a shift of particle number mode diameter to larger ones. For experiments α-pinene ozonolysis and γ-terpinene ozonolysis. While NOx obviously
with initial ~15 ppb γ-terpinene, the number mode diameter of particles enhance γ-terpinene SOA formation over the [γ-terpinene]0/[NOx]0
increased from 131 nm in pure ozonolysis reaction to 163 nm when range studied, previous studies suggested that increasing NOx concen­
[γ-terpinene]0/[NOx]0 ratio reached 5.0. A similar effect of NOx on the tration substantially deplete SOA formation from α-pinene ozonolysis
particle size distribution was also observed from the ozonolysis of (Draper et al., 2015; Presto et al., 2005b). Such a negative NOx effect on
α-pinene (~50 ppb) and d-limonene (~50 ppb). The presence of 50 ppb α-pinene ozonolysis was attributed to high volatile products and near
NO2 leads the particle mode diameter to increase by 20 nm for α-pinene

5
L. Xu et al. Atmospheric Environment 225 (2020) 117375

Fig. 4. SOA yields from NOx-involved γ-terpinene ozonolysis (Exp. 9-24). Fig. 5. Time series of γ-terpinene consumption by O3 or NO3 under different
NOx conditions (Exp. 11 and Exp.14).

zero SOA yield from NO3 oxidation (Fry et al., 2014). The effects of NOx
on SOA formation from the ozonolysis of β-pinene, Δ3-carene and making the γ-terpinene consumption to shift from O3-dominated to
limonene have been investigated using a dark flow through a chamber NO3-dominated oxidation. Thereby, the particle size and SOA yields
by Draper et al. and it was found that by keeping O3 mixing ratio con­ increase can be caused by the increased NO3 radical contribution as
stant and varying NOx mixing ratio, SOA yields from β-pinene and expected (Figs. 3 and 4).
Δ3-carene ozonolysis were comparable over the range of oxidation
conditions ([O3]/[NO2] ¼ 2–0.5, [NO2]/[VOC] ¼ 0.5–1) while limo­
3.3. Effect of NOx on SOA composition
nene SOA yield increased with NOx concentration (Draper et al., 2015).
The later observation was explained by increased multifunctional
NO3-induced BVOC oxidation is a potentially efficient pathway for
organic nitrates and oligomers fraction in SOA at higher NOx.
the formation of organic nitrates, which can partition into the condensed
The increased SOA yields with NOx mixing ratios may result from the
phase and contribute to SOA formation and growth (Faxon et al., 2018;
enhanced competition between NO3- and O3-initiated γ-terpinene
Ng et al., 2017). In the presence of NOx, NO3 radical formation and its
oxidation. To quantify the relative contributions of NO3 radical and O3
further competition with O3 are expected to produce particle-phase
to γ-terpinene consumption under different NOx conditions, we built up
organic nitrates. To get detailed information about functional groups
a simple kinetic box model (Draper et al., 2015; Fry et al., 2009; Noj­
in particle-phase products, SOA collected onto aluminum foil pieces
gaard et al., 2006). The concentration evolution of γ-terpinene and its
(Exp. 16-24) were analyzed by ATR-FTIR. Some typical infrared spectra
loss ascribed to the O3/NO3 reaction were estimated. The model was
of SOA generated from three NOx level experiments (NOx-free, [γ-ter­
initialized with the observed O3, NOx and the mass of γ-terpinene
pinene]0/NOx]0 ¼ 20.8, and [γ-terpinene]0/NOx]0 ¼ 3.9) are shown in
injected and constrained with the known reaction rates. Reactions and
Fig. 6. All SOA samples exhibit strong absorption at 1713 cm 1, which is
corresponding rate constants incorporated in this model are summarized
attributed to the stretching vibration of the carbonyl moiety (Allen et al.,
in Table S2. The rate constant of the gas-phase reaction between O3 and
1994; Presto et al., 2005b). When NOx were present in the γ-terpinene
NO2 at 298 K is 3.2 � 10 17 cm3 molec 1 s 1, about four times lower
ozonolysis system, three new strong absorption peaks appeared at 1635,
than the rate constant (1.4 � 10 16 cm3 molec 1 s 1) of the reaction
between O3 and γ-terpinene (Atkinson et al., 1990; Sander et al., 2011).
This shows that the NO3 radical formation may be not favored until the
mixing ratio of NOx is about four times higher than that of γ-terpinene
during the reaction process. The initial consumption of γ-terpinene is
dominated by ozonolysis, especially in experiments with lower NOx
mixing ratios. As modeled results show (Fig. 5), O3 is the dominant
oxidant when the initial NOx mixing ratio is 76 ppb ([γ-terpinene]0/­
NOx]0 ¼ 20.3). When NOx concentration is elevated to 310 ppb
([γ-terpinene]0/NOx]0 ¼ 5.0), the competition between O3 and NO3
radical for γ-terpinene shifts from O3-dominated to NO3-dominated
oxidation at ~10 min after the initiation of the reaction. The SOA yield
at this [γ-terpinene]0/NOx]0 ratio is 0.67, 0.29 higher than that in pure
ozonolysis (Exp.5). In a previous study, the homogeneous nucleation
potential of monoterpenes, especially endocyclic monoterpenes, was
suggested to be much higher during ozonolysis than in oxidations by
NO3 or OH radicals (Bonn and Moorgat, 2002). The initial consumption
of γ-terpinene by O3 is thus expected to be primarily responsible for new
particle formation. The products from oxidations by NO3 radical and O3
would be both incorporated into SOA growth through condensation
(Perraud et al., 2012). As NOx concentration increases, the decreased
γ-terpinene consumption by O3 results in more reactant for NO3 radical, Fig. 6. Infrared spectra of SOA generated under different NOx conditions (Exp.
16, 19, 24).

6
L. Xu et al. Atmospheric Environment 225 (2020) 117375

1281 and 854 cm 1, corresponding to the asymmetric stretching of NO2, elevated NOx levels.
symmetric stretching of NO2, and symmetric stretching of NO, respec­ The SOA composition from pure and NOx-involved γ-terpinene
tively (Bruns et al., 2010; Hallquist et al., 1999). Besides these three ozonolysis (Exp. 25 and 26) was further examined by UPLC/ESI-HR-Q-
prominent organonitrate peaks, two other weaker absorptions at 696 TOFMS operated in the negative ionization mode. Due to the presence
and 752 cm 1 could also be assigned to the characteristic absorption of of ozone in NOx-free and NOx-involved experiments, the ozonolysis of
organic nitrates according to the infrared spectra of 2-ethylhexyl nitrate γ-terpinene occurred in both cases. As shown in Fig. 8, most commonly
(Day et al., 2010). The wide absorption shoulder from 985 to 1243 cm 1 shared mass spectrum peaks could be the products from γ-terpinene
is assigned to C–O and C–O–O groups while the smaller absorption at ozonolysis. In the pure ozonolysis of γ-terpinene, where O3 is the only
957 cm 1 matches the O–O group (Jia and Xu, 2018). The coexistence of oxidant, four Criegee intermediates (CIs) are likely to form considering
multiple functional groups in the infrared spectrum illustrates that similar character of the two double bonds (Fig. S2). The subsequent
multifunctional organic nitrates may be an important component of reactions of these CIs may be responsible for the species detected in the
SOA. particle-phase. Based on the accurate m/z values in the negative ioni­
The infrared spectra of SOA from γ-terpinene ozonolysis are char­ zation mass spectra, compounds in SOA that can be formed through CIs
acterized by the dominated absorption of carbonyl group, which grad­ reactions are tentatively identified and the possible formation channels
ually lost its superiority as NOx levels were elevated. To further evaluate are proposed (Fig. 9). Unimolecular reactions of excited CIs lead to
the trends of organic nitrates production with varying initial NOx con­ various products, including organic acids, aldehydes, and carbonyls
ditions, infrared peak heights of organic nitrates (1635, 1281, 854 and (Tobias and Ziemann, 2001). Most of these species are in the mid mass
752 cm 1) are normalized with respect to the carbonyl group at 1713 range of 160 < m/z < 250. Products with larger molecular weight (m/z
cm 1 (Fig. 7). Clearly, the relative peak heights for organic nitrates in­ > 250) in the particle phase can be formed via bimolecular reactions of
crease with the increasing NOx mixing ratios. It is estimated that adding collisionally stabilized CIs with organic acids, carbonyls or some other
organic nitrate functional groups to hydrocarbons can reduce the equi­ reactions (e.g. hemi-acetal and aldol condensations) (Heaton et al.,
librium saturated vapor pressure by 2–3 orders of magnitude (Lee et al., 2007; Kundu et al., 2012; Mackenzie-Rae et al., 2018). For example, m/z
2016). The increased organic nitrates fraction in SOA samples may 313 and 329 product ions are detected in the SOA sample with high
partially explain the dramatic increase of SOA yields at higher NOx relative intensity (>8%). Corresponding elemental compositions are
conditions (Fig. 4). Under lower NOx levels, for example, [γ-terpine­ identified as C16H26O6 and C16H26O7 (ΔM/M < 10 5), respectively.
ne]0/[NOx]0 ¼ 51. (Exp.18), the normalized peak heights for organic These two secondary ozonides can be formed through the reaction of
nitrates at 1635, 1230, 854 cm 1 are 0.58, 0.65, and 0.35, respectively, stabilized CIs with carbonyl compounds (Heaton et al., 2007). Due to the
lower than the carbonyl absorbance at 1713 cm 1. The SOA yield at this presence of two double bonds in γ-terpinene, the ozonolysis of the sec­
NOx mixing ratio is 0.17 higher than that from pure ozonolysis (Exp.16). ond double bond produces species with smaller molecular weight. Some
Peak height ratios of organic nitrates to the carbonyl, namely of these second-generation products can partition into the particle-phase
1635/1713, 1230/1713, 854/1713, reach 1.8, 1.6 and 1.3 when as observed in the mass spectra. For example, products ions of m/z 127,
[γ-terpinene]0/[NOx]0 decreases to 3.9 (Exp.24). Compared with pure 129,145 can be formed via the subsequent reaction of a C6 CI that is
ozonolysis, the SOA yield increases by 0.75. Enhanced organic nitrates generated by the ozonolysis of first-generation products. Overall, the
formation at higher NOx is consistent with NO3 radical formation. nucleation, absorptive partitioning of ozonolysis products between the
Combining with increased SOA yields, products containing the organic gas and particle phases and coagulation promote γ-terpinene SOA for­
nitrate moiety may be less volatile than species formed from NOx-free mation and growth.
conditions. Organic nitrates were also observed to be important SOA In NOx-involved γ-terpinene ozonolysis, some new peaks appear
components from limonene ozonolysis. With varying NO2 concentra­ with relatively high intensities (Fig. 8). NOx affect the SOA composition
tions, organic nitrates produced by NO3-induced limonene oxidation in two different channels as discussed above. NO3 chemistry is expected
accounted for about 30% of the total SOA mass (Draper et al., 2015; Fry to be the efficient way to influence SOA composition (Draper et al.,
et al., 2011). Organic nitrates in γ-terpinene SOA indicate that γ-terpi­ 2015). Following NO3 radical oxidation mechanism, some new peaks are
nene is also an important source of organic nitrates in regions with tentatively identified as organic nitrates based on the accurate mass
(ΔM/M < 10 5) and their absence in the pure-ozonolysis.

Fig. 7. Peak height ratio of organic nitrates (1635, 1280, 854, 752 cm 1) to Fig. 8. The ESI mass spectrum (negative ionization mode) of SOA form pure
carbonyl (1713 cm 1) for SOA generated under different NOx conditions (Exp. (Exp. 25) and NOx-involved (Exp. 26) γ-terpinene ozonolysis. Ion peaks with
16, 18, 19, 21, 24). relative intensity greater than 3% of the strongest peak are shown.

7
L. Xu et al. Atmospheric Environment 225 (2020) 117375

Fig. 9. Proposed formation mechanisms for compounds in SOA from γ-terpinene ozonolysis (pink structures). The pink structures are tentatively identified based on
the negative ionization mode ESI mass spectrum from UPLC/ESI-HR-Q-TOFMS detection. The partial formation pathway for high molecular weight products (m/z ¼
313, 329) via the reaction between stabilized CIs and carbonyl is shown in the dotted box. The formation of some second-generation products from the ozonolysis of
the first-generation products (m/z ¼ 127, 129, 145) is shown in the solid box. Other reaction pathways and isomers are possible but are not shown here for clarity.
(For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

Corresponding reaction pathways are proposed in Fig. 10. Note that could be well fitted by a two-product model. The relatively low SOA
ozone and NO3 radical are simultaneously present as oxidants in the yield under atmospherically relevant SOA mass loadings suggests that
reaction system with NOx, first-generation products from γ-terpinene γ-terpinene may not be an important SOA precursor in regions that are
oxidation by NO3 radical could be further oxidized by ozone to form less affected by NOx emissions.
second-generation products. The observation of most of new peaks (e.g., NOx effects on γ-terpinene ozonolysis were studied with varying NOx
m/z ¼ 278, 346, 266, 262) could be explained by these conditions. The [NOx]0/[O3]0 ratio ranged approximately from 0 to 0.7
second-generation multifunctional organic nitrates. The dominant new and [γ-terpinene]0/[NOx]0 varied in a relatively larger range ([γ-terpi­
peaks are in the region m/z > 250, which comprises a greater fraction of nene]0/[NOx]0 ¼ 4-100). These conditions can be representative of
total ion intensity than in pure ozonolysis. The organic nitrates identi­ relative clean or NOx polluted night atmospheric environments in some
fied here are in line with the strong absorbance of organic nitrates in the regions (Chen et al., 2019b; Millet et al., 2016). When different levels of
infrared spectrum. Hence, the most important pathway for NOx to affect NOx were considered in the γ-terpinene ozonolysis system, we found
SOA yields and composition is to be converted to NO3 radical that can that this process produced particles with larger mode diameter at rela­
oxidize γ-terpinene directly. Both the first and subsequent tively higher NOx levels. Meanwhile, SOA yields increased markedly
second-generation products contribute substantially to SOA growth and with the decreasing [γ-terpinene]0/[NOx]0 ratios. The SOA composition
SOA yield. analysis by ATR-FTIR showed that the fraction of organic nitrates in­
creases with increasing NOx concentrations. The characterization of
4. Conclusions chemical components by UPLC/ESI-HR-Q-TOFMS and the mechanism
analysis suggested that multifunctional organic nitrates with high mo­
The nighttime atmospheric oxidation capacity is highly relevant to lecular mass (m/z > 250) can be formed in NOx-involved γ-terpinene
NOx levels. In relatively clean regions that are isolated from NOx ozonolysis.
emission sources, O3 is the dominant oxidant for BVOCs oxidation. The enhancement effect of NOx on SOA yields and organic nitrates
When NOx are transported to densely forested areas with high BVOC formation can be mainly explained by the generation of NO3 radical and
emissions or vice versa, the dominance of O3 for BVOC oxidation can be its competition with O3. The first-generation products from γ-terpinene
altered (Chen et al., 2019b). In the present study, nighttime SOA for­ oxidation by NO3 radical could be further oxidized by O3, forming
mation from γ-terpinene ozonolysis and NOx effects on this process were multifunctional organic nitrates that can contribute to SOA formation.
both studied under dark conditions. With varying initial γ-terpinene In regions with high NOx levels, the more favored γ-terpinene oxidation
mixing ratios, a set of SOA yields (0.14–0.55) were obtained in the by NO3 radical is likely to greatly accelerate γ-terpinene consumption,
aerosol mass loading ranging from 25 to 660 μg m 3. The yield data limit its atmospheric lifetime and most importantly, strengthen organic

8
L. Xu et al. Atmospheric Environment 225 (2020) 117375

Fig. 10. Proposed formation mechanisms for compounds in SOA generated in NOx-involved γ-terpinene ozonolysis. The pink structures are tentatively identified as
organic nitrates based on the negative ionization mode ESI mass spectrum from UPLC/ESI-HR-Q-TOFMS detection. (For interpretation of the references to colour in
this figure legend, the reader is referred to the Web version of this article.)

nitrates fraction in SOA. Acknowledgements


Organic nitrates contribute substantially to nighttime organic aero­
sol as many recent field observations suggested (Kiendler-Scharr et al., This work was supported by National Natural Science Foundation of
2016; Sobanski et al., 2017; Yu et al., 2019). Our study corroborates the China (91644214), Shandong Natural Science Fund for Distinguished
important role of NOx on nighttime γ-terpenene oxidation, suggests also Young Scholars (JQ201705), and Fundamental Research Fund of
that γ-terpinene is an important BVOC precursor that can contribute to Shandong University (2020QNQT012).
particle-phase organic nitrates and SOA formation in the nighttime.
Models that are used to predict regional or global SOA mass loadings of Appendix A. Supplementary data
organic nitrates formation should take into account γ-terpinene oxida­
tion, especially in regions influenced by both anthropogenic pollutants Supplementary data to this article can be found online at https://doi.
and high γ-terpinene emissions. org/10.1016/j.atmosenv.2020.117375.

Declaration of competing interest References

The authors declare that they have no known competing financial Allen, D.T., Palen, E.J., Haimov, M.I., Hering, S.V., Young, J.R., 1994. Fourier transform
interests or personal relationships that could have appeared to influence infrared spectroscopy of aerosol collected in a low-pressure impactor (LPI/FTIR) -
method development and field calibration. Aerosol Sci. Technol. 21, 325–342.
the work reported in this paper. Atkinson, R., 2007. Gas-phase tropospheric chemistry of organic compounds: a review.
Atmos. Environ. 41, 200–240.
Atkinson, R., Hasegawa, D., Aschmann, S.M., 1990. Rate constants for the gas-phase
CRediT authorship contribution statement reactions of O3 with a series of monoterpenes and related compounds at 296�2 K.
Int. J. Chem. Kinet. 22, 871–887.
Li Xu: Conceptualization, Methodology, Investigation, Writing - Berkemeier, T., Ammann, M., Mentel, T.F., Poschl, U., Shiraiwa, M., 2016. Organic
nitrate contribution to new particle formation and growth in secondary organic
original draft. Narcisse T. Tsona: Writing - review & editing, Visuali­
aerosols from α-pinene ozonolysis. Environ. Sci. Technol. 50, 6334–6342.
zation. Bo You: Investigation. Yingnan Zhang: Methodology, Software. Bonn, B., Moorgat, G., 2002. New particle formation during α-and β-pinene oxidation by
Shuyan Wang: Investigation. Zhaomin Yang: Investigation. Likun O3, OH and NO3, and the influence of water vapour: particle size distribution studies.
Xue: Methodology, Software. Lin Du: Conceptualization, Funding Atmos. Chem. Phys. 2, 183–196.
Boyd, C.M., Sanchez, J., Xu, L., Eugene, A.J., Nah, T., Tuet, W.Y., Guzman, M.I., Ng, N.L.,
acquisition, Project administration, Resources, Supervision, 2015. Secondary organic aerosol formation from the β-pineneþNO3 system: effect of
Visualization. humidity and peroxy radical fate. Atmos. Chem. Phys. 15, 7497–7522.

9
L. Xu et al. Atmospheric Environment 225 (2020) 117375

Brown, S.S., Stutz, J., 2012. Nighttime radical observations and chemistry. Chem. Soc. Jimenez, J.L., Canagaratna, M.R., Donahue, N.M., Prevot, A.S., Zhang, Q., Kroll, J.H.,
Rev. 41, 6405–6447. DeCarlo, P.F., Allan, J.D., Coe, H., Ng, N.L., Aiken, A.C., Docherty, K.S., Ulbrich, I.
Bruns, E.A., Perraud, V., Zelenyuk, A., Ezell, M.J., Johnson, S.N., Yu, Y., Imre, D., M., Grieshop, A.P., Robinson, A.L., Duplissy, J., Smith, J.D., Wilson, K.R., Lanz, V.A.,
Finlayson-Pitts, B.J., Alexander, M.L., 2010. Comparison of FTIR and particle mass Hueglin, C., Sun, Y.L., Tian, J., Laaksonen, A., Raatikainen, T., Rautiainen, J.,
spectrometry for the measurement of particulate organic nitrates. Environ. Sci. Vaattovaara, P., Ehn, M., Kulmala, M., Tomlinson, J.M., Collins, D.R., Cubison, M.J.,
Technol. 44, 1056–1061. Dunlea, E.J., Huffman, J.A., Onasch, T.B., Alfarra, M.R., Williams, P.I., Bower, K.,
Budisulistiorini, S., Li, X., Bairai, S., Renfro, J., Liu, Y., Liu, Y., McKinney, K., Martin, S., Kondo, Y., Schneider, J., Drewnick, F., Borrmann, S., Weimer, S., Demerjian, K.,
McNeill, V., Pye, H., 2015. Examining the effects of anthropogenic emissions on Salcedo, D., Cottrell, L., Griffin, R., Takami, A., Miyoshi, T., Hatakeyama, S.,
isoprene-derived secondary organic aerosol formation during the 2013 Southern Shimono, A., Sun, J.Y., Zhang, Y.M., Dzepina, K., Kimmel, J.R., Sueper, D., Jayne, J.
Oxidant and Aerosol Study (SOAS) at the Look Rock, Tennessee ground site. Atmos. T., Herndon, S.C., Trimborn, A.M., Williams, L.R., Wood, E.C., Middlebrook, A.M.,
Chem. Phys. 15, 8871–8888. Kolb, C.E., Baltensperger, U., Worsnop, D.R., 2009. Evolution of organic aerosols in
Chen, T.Z., Liu, Y.C., Ma, Q.X., Chu, B.W., Zhang, P., Liu, C.G., Liu, J., He, H., 2019a. the atmosphere. Science 326, 1525–1529.
Significant source of secondary aerosol: formation from gasoline evaporative Jokinen, T., Berndt, T., Makkonen, R., Kerminen, V.M., Junninen, H., Paasonen, P.,
emissions in the presence of SO2 and NH3. Atmos. Chem. Phys. 19, 8063–8081. Stratmann, F., Herrmann, H., Guenther, A.B., Worsnop, D.R., Kulmala, M., Ehn, M.,
Chen, X.R., Wang, H.C., Liu, Y.H., Su, R., Wang, H.L., Lou, S.R., Lu, K.D., 2019b. Spatial Sipila, M., 2015. Production of extremely low volatile organic compounds from
characteristics of the nighttime oxidation capacity in the Yangtze River Delta, China. biogenic emissions: measured yields and atmospheric implications. Proc. Natl. Acad.
Atmos. Environ. 208, 150–157. Sci. U.S.A. 112, 7123–7128.
Chu, B., Zhang, X., Liu, Y., He, H., Sun, Y., Jiang, J., Li, J., Hao, J., 2016. Synergetic Kavouras, I.G., Mihalopoulos, N., Stephanou, E.G., 1998. Formation of atmospheric
formation of secondary inorganic and organic aerosol: effect of SO2 and NH3 on particles from organic acids produced by forests. Nature 395, 683–686.
particle formation and growth. Atmos. Chem. Phys. 16, 14219–14230. Kiendler-Scharr, A., Mensah, A.A., Friese, E., Topping, D., Nemitz, E., Prevot, A.S.H.,
Day, D.A., Liu, S., Russell, L.M., Ziemann, P.J., 2010. Organonitrate group concentrations Aij€
€ al€a, M., Allan, J., Canonaco, F., Canagaratna, M., Carbone, S., Crippa, M., Dall
in submicron particles with high nitrate and organic fractions in coastal southern Osto, M., Day, D.A., De Carlo, P., Di Marco, C.F., Elbern, H., Eriksson, A., Freney, E.,
California. Atmos. Environ. 44, 1970–1979. Hao, L., Herrmann, H., Hildebrandt, L., Hillamo, R., Jimenez, J.L., Laaksonen, A.,
Donahue, N.M., Ortega, I.K., Chuang, W., Riipinen, I., Riccobono, F., Schobesberger, S., McFiggans, G., Mohr, C., O’Dowd, C., Otjes, R., Ovadnevaite, J., Pandis, S.N.,
Dommen, J., Baltensperger, U., Kulmala, M., Worsnop, D.R., 2013. How do organic Poulain, L., Schlag, P., Sellegri, K., Swietlicki, E., Tiitta, P., Vermeulen, A.,
vapors contribute to new-particle formation? Faraday Discuss 165, 91–104. Wahner, A., Worsnop, D., Wu, H.C., 2016. Ubiquity of organic nitrates from
Draper, D.C., Farmer, D.K., Desyaterik, Y., Fry, J.L., 2015. A qualitative comparison of nighttime chemistry in the European submicron aerosol. Geophys. Res. Lett. 43,
secondary organic aerosol yields and composition from ozonolysis of monoterpenes 7735–7744.
at varying concentrations of NO2. Atmos. Chem. Phys. 15, 12267–12281. Kourtchev, I., Doussin, J.F., Giorio, C., Mahon, B., Wilson, E.M., Maurin, N., Pangui, E.,
Ehn, M., Thornton, J.A., Kleist, E., Sipila, M., Junninen, H., Pullinen, I., Springer, M., Venables, D.S., Wenger, J.C., Kalberer, M., 2015. Molecular composition of fresh and
Rubach, F., Tillmann, R., Lee, B., Lopez-Hilfiker, F., Andres, S., Acir, I.H., aged secondary organic aerosol from a mixture of biogenic volatile compounds: a
Rissanen, M., Jokinen, T., Schobesberger, S., Kangasluoma, J., Kontkanen, J., high-resolution mass spectrometry study. Atmos. Chem. Phys. 15, 5683–5695.
Nieminen, T., Kurten, T., Nielsen, L.B., Jorgensen, S., Kjaergaard, H.G., Krechmer, J.E., Pagonis, D., Ziemann, P.J., Jimenez, J.L., 2016. Quantification of gas-
Canagaratna, M., Dal Maso, M., Berndt, T., Petaja, T., Wahner, A., Kerminen, V.M., wall partitioning in Teflon environmental chambers using rapid bursts of low-
Kulmala, M., Worsnop, D.R., Wildt, J., Mentel, T.F., 2014. A large source of low- volatility oxidized species generated in situ. Environ. Sci. Technol. 50, 5757–5765.
volatility secondary organic aerosol. Nature 506, 476–479. Kundu, S., Fisseha, R., Putman, A.L., Rahn, T.A., Mazzoleni, L.R., 2012. High molecular
Esquivel-Hern� andez, G., Madrigal-Carballo, S., Alfaro-Solís, R., Sibaja-Brenes, J.P., weight SOA formation during limonene ozonolysis: insights from ultrahigh-
Vald�es-Gonz� alez, J., 2011. First measurements of biogenic hydrocarbons in air in a resolution FT-ICR mass spectrometry characterization. Atmos. Chem. Phys. 12,
tropical cloudy forest, Monteverde, Costa Rica. J. Chem. Chem. Eng. 1097–1106. 5523–5536.
Faxon, C., Hammes, J., Le Breton, M., Pathak, R.K., Hallquist, M., 2018. Characterization Lane, T.E., Donahue, N.M., Pandis, S.N., 2008. Effect of NOx on secondary organic
of organic nitrate constituents of secondary organic aerosol (SOA) from nitrate- aerosol concentrations. Environ. Sci. Technol. 42, 6022–6027.
radical-initiated oxidation of limonene using high-resolution chemical ionization Lee, A., Goldstein, A.H., Keywood, M.D., Gao, S., Varutbangkul, V., Bahreini, R., Ng, N.
mass spectrometry. Atmos. Chem. Phys. 18, 5467–5481. L., Flagan, R.C., Seinfeld, J.H., 2006. Gas-phase products and secondary aerosol
Friedman, B., Farmer, D.K., 2018. SOA and gas phase organic acid yields from the yields from the ozonolysis of ten different terpenes. J. Geophys. Res. Atmos. 111,
sequential photooxidation of seven monoterpenes. Atmos. Environ. 187, 335–345. D07302.
Fry, J.L., Draper, D.C., Barsanti, K.C., Smith, J.N., Ortega, J., Winkler, P.M., Lawler, M.J., Lee, B.H., Mohr, C., Lopez-Hilfiker, F.D., Lutz, A., Hallquist, M., Lee, L., Romer, P.,
Brown, S.S., Edwards, P.M., Cohen, R.C., Lee, L., 2014. Secondary organic aerosol Cohen, R.C., Iyer, S., Kurten, T., Hu, W.W., Day, D.A., Campuzano-Jost, P.,
formation and organic nitrate yield from NO3 oxidation of biogenic hydrocarbons. Jimenez, J.L., Xu, L., Ng, N.L., Guo, H.Y., Weber, R.J., Wild, R.J., Brown, S.S.,
Environ. Sci. Technol. 48, 11944–11953. Koss, A., de Gouw, J., Olson, K., Goldstein, A.H., Seco, R., Kim, S., McAvey, K.,
Fry, J.L., Kiendler-Scharr, A., Rollins, A.W., Brauers, T., Brown, S.S., Dorn, H.P., Shepson, P.B., Starn, T., Baumann, K., Edgerton, E.S., Liu, J.M., Shilling, J.E.,
Dube, W.P., Fuchs, H., Mensah, A., Rohrer, F., Tillmann, R., Wahner, A., Miller, D.O., Brune, W., Schobesberger, S., D’Ambro, E.L., Thornton, J.A., 2016.
Wooldridge, P.J., Cohen, R.C., 2011. SOA from limonene: role of NO3 in its Highly functionalized organic nitrates in the southeast United States: contribution to
generation and degradation. Atmos. Chem. Phys. 11, 3879–3894. secondary organic aerosol and reactive nitrogen budgets. Proc. Natl. Acad. Sci. U.S.
Fry, J.L., Kiendler-Scharr, A., Rollins, A.W., Wooldridge, P.J., Brown, S.S., Fuchs, H., A. 113, 1516–1521.
Dube, W., Mensah, A., dal Maso, M., Tillmann, R., Dorn, H.P., Brauers, T., Cohen, R. Liu, J., Russell, L.M., Ruggeri, G., Takahama, S., Claflin, M.S., Ziemann, P.J., Pye, H.O.,
C., 2009. Organic nitrate and secondary organic aerosol yield from NO3 oxidation of Murphy, B.N., Xu, L., Ng, N.L., 2018. Regional similarities and NOx-related increases
β-pinene evaluated using a gas-phase kinetics/aerosol partitioning model. Atmos. in biogenic secondary organic aerosol in summertime southeastern United States.
Chem. Phys. 9, 1431–1449. J. Geophys. Res. Atmos. 123, 10,620–10,636.
Geron, C., Rasmussen, R., Arnts, R.R., Guenther, A., 2000. A review and synthesis of Mackenzie-Rae, F.A., Wallis, H.J., Rickard, A.R., Pereira, K.L., Saunders, S.M., Wang, X.,
monoterpene speciation from forests in the United States. Atmos. Environ. 34, Hamilton, J.F., 2018. Ozonolysis of α-phellandrene-Part 2: compositional analysis of
1761–1781. secondary organic aerosol highlights the role of stabilized Criegee intermediates.
Griffin, R.J., Cocker, D.R., Flagan, R.C., Seinfeld, J.H., 1999. Organic aerosol formation Atmos. Chem. Phys. 4673–4693.
from the oxidation of biogenic hydrocarbons. J. Geophys. Res. Atmos. 104, Matsui, H., Koike, M., Kondo, Y., Takami, A., Fast, J.D., Kanaya, Y., Takigawa, M., 2014.
3555–3567. Volatility basis-set approach simulation of organic aerosol formation in East Asia:
Guenther, A., Hewitt, C.N., Erickson, D., Fall, R., Geron, C., Graedel, T., Harley, P., implications for anthropogenic–biogenic interaction and controllable amounts.
Klinger, L., Lerdau, M., McKay, W.A., Pierce, T., Scholes, B., Steinbrecher, R., Atmos. Chem. Phys. 14, 9513–9535.
Tallamraju, R., Taylor, J., Zimmerman, P., 1995. A global-model of natural volatile Mauderly, J.L., Chow, J.C., 2008. Health effects of organic aerosols. Inhal. Toxicol. 20,
organic-compound emissions. J. Geophys. Res. Atmos. 100, 8873–8892. 257–288.
Hallquist, M., Wangberg, I., Ljungstrom, E., Barnes, I., Becker, K.H., 1999. Aerosol and Millet, D.B., Baasandorj, M., Hu, L., Mitroo, D., Turner, J., Williams, B.J., 2016.
product yields from NO3 radical-initiated oxidation of selected monoterpenes. Nighttime chemistry and morning isoprene can drive urban ozone downwind of a
Environ. Sci. Technol. 33, 553–559. major deciduous forest. Environ. Sci. Technol. 50, 4335–4342.
Hallquist, M., Wenger, J.C., Baltensperger, U., Rudich, Y., Simpson, D., Claeys, M., Moldanova, J., Ljungstr€ om, E., 2000. Modelling of particle formation from NO3 oxidation
Dommen, J., Donahue, N.M., George, C., Goldstein, A.H., Hamilton, J.F., of selected monoterpenes. J. Aerosol Sci. 31, 1317–1333.
Herrmann, H., Hoffmann, T., Iinuma, Y., Jang, M., Jenkin, M.E., Jimenez, J.L., Moortgat, B.B.a.G.K., 2002. New particle formation during α- and β-pinene oxidation by
Kiendler-Scharr, A., Maenhaut, W., McFiggans, G., Mentel, T.F., Monod, A., O3, OH and NO3, and the influence of water vapour: particle size distribution studies.
Prevot, A.S.H., Seinfeld, J.H., Surratt, J.D., Szmigielski, R., Wildt, J., 2009. The Atmos. Chem. Phys. 2, 183–196.
formation, properties and impact of secondary organic aerosol: current and Nah, T., McVay, R.C., Zhang, X., Boyd, C.M., Seinfeld, J.H., Ng, N.L., 2016. Influence of
emerging issues. Atmos. Chem. Phys. 9, 5155–5236. seed aerosol surface area and oxidation rate on vapor wall deposition and SOA mass
Han, Y.M., Stroud, C.A., Liggio, J., Li, S.M., 2016. The effect of particle acidity on yields: a case study with α-pinene ozonolysis. Atmos. Chem. Phys. 16, 9361–9379.
secondary organic aerosol formation from α-pinene photooxidation under Ng, N.L., Brown, S.S., Archibald, A.T., Atlas, E., Cohen, R.C., Crowley, J.N., Day, D.A.,
atmospherically relevant conditions. Atmos. Chem. Phys. 16, 13929–13944. Donahue, N.M., Fry, J.L., Fuchs, H., Griffin, R.J., Guzman, M.I., Herrmann, H.,
Heaton, K.J., Dreyfus, M.A., Wang, S., Johnston, M.V., 2007. Oligomers in the early stage Hodzic, A., Iinuma, Y., Jimenez, J.L., Kiendler-Scharr, A., Lee, B.H., Luecken, D.J.,
of biogenic secondary organic aerosol formation and growth. Environ. Sci. Technol. Mao, J., McLaren, R., Mutzel, A., Osthoff, H.D., Ouyang, B., Picquet-Varrault, B.,
41, 6129–6136. Platt, U., Pye, H.O.T., Rudich, Y., Schwantes, R.H., Shiraiwa, M., Stutz, J.,
Jia, L., Xu, Y., 2018. Different roles of water in secondary organic aerosol formation from Thornton, J.A., Tilgner, A., Williams, B.J., Zaveri, R.A., 2017. Nitrate radicals and
toluene and isoprene. Atmos. Chem. Phys. 18, 8137–8154.

10
L. Xu et al. Atmospheric Environment 225 (2020) 117375

biogenic volatile organic compounds: oxidation, mechanisms, and organic aerosol. Sobanski, N., Thieser, J., Schuladen, J., Sauvage, C., Song, W., Williams, J., Lelieveld, J.,
Atmos. Chem. Phys. 17, 2103–2162. Crowley, J.N., 2017. Day and night-time formation of organic nitrates at a forested
Nojgaard, J.K., Bilde, M., Stenby, C., Nielsen, O.J., Wolkoff, P., 2006. The effect of mountain site in south-west Germany. Atmos. Chem. Phys. 17, 4115–4130.
nitrogen dioxide on particle formation during ozonolysis of two abundant Song, C., Zaveri, R.A., Alexander, M.L., Thornton, J.A., Madronich, S., Ortega, J.V.,
monoterpenes indoors. Atmos. Environ. 40, 1030–1042. Zelenyuk, A., Yu, X.Y., Laskin, A., Maughan, D.A., 2007. Effect of hydrophobic
Odum, J.R., Hoffmann, T., Bowman, F., Collins, D., Flagan, R.C., Seinfeld, J.H., 1996. primary organic aerosols on secondary organic aerosol formation from ozonolysis of
Gas/particle partitioning and secondary organic aerosol yields. Environ. Sci. α-pinene. Geophys. Res. Lett. 34, L20803.
Technol. 30, 2580–2585. Spracklen, D.V., Jimenez, J.L., Carslaw, K.S., Worsnop, D.R., Evans, M.J., Mann, G.W.,
Perraud, V., Bruns, E.A., Ezell, M.J., Johnson, S.N., Greaves, J., Finlayson-Pitts, B.J., Zhang, Q., Canagaratna, M.R., Allan, J., Coe, H., McFiggans, G., Rap, A., Forster, P.,
2010. Identification of organic nitrates in the NO3 radical initiated oxidation of 2011. Aerosol mass spectrometer constraint on the global secondary organic aerosol
α-pinene by atmospheric pressure chemical ionization mass spectrometry. Environ. budget. Atmos. Chem. Phys. 11, 12109–12136.
Sci. Technol. 44, 5887–5893. Stirnweis, L., Marcolli, C., Dommen, J., Barmet, P., Frege, C., Platt, S.M., Bruns, E.A.,
Perraud, V., Bruns, E.A., Ezell, M.J., Johnson, S.N., Yu, Y., Alexander, M.L., Zelenyuk, A., Krapf, M., Slowik, J.G., Wolf, R., Pr�ev^ot, A.S.H., Baltensperger, U., El-Haddad, I.,
Imre, D., Chang, W.L., Dabdub, D., Pankow, J.F., Finlayson-Pitts, B.J., 2012. 2017. Assessing the influence of NOx concentrations and relative humidity on
Nonequilibrium atmospheric secondary organic aerosol formation and growth. Proc. secondary organic aerosol yields from α-pinene photo-oxidation through smog
Natl. Acad. Sci. U.S.A. 109, 2836–2841. chamber experiments and modelling calculations. Atmos. Chem. Phys. 17,
Pratt, K.A., Mielke, L.H., Shepson, P.B., Bryan, A.M., Steiner, A.L., Ortega, J., Daly, R., 5035–5061.
Helmig, D., Vogel, C.S., Griffith, S., Dusanter, S., Stevens, P.S., Alaghmand, M., 2012. Takekawa, H., Minoura, H., Yamazaki, S., 2003. Temperature dependence of secondary
Contributions of individual reactive biogenic volatile organic compounds to organic organic aerosol formation by photo-oxidation of hydrocarbons. Atmos. Environ. 37,
nitrates above a mixed forest. Atmos. Chem. Phys. 12, 10125–10143. 3413–3424.
Presto, A.A., Hartz, K.E.H., Donahue, N.M., 2005a. Secondary organic aerosol production Tobias, H.J., Ziemann, P.J., 2001. Kinetics of the gas-phase reactions of alcohols,
from terpene ozonolysis. 1. Effect of UV radiation. Environ. Sci. Technol. 39, aldehydes, carboxylic acids, and water with the C13 stabilized Criegee intermediate
7036–7045. formed from ozonolysis of 1-tetradecene. J. Phys. Chem. A 105, 6129–6135.
Presto, A.A., Hartz, K.E.H., Donahue, N.M., 2005b. Secondary organic aerosol production Virtanen, A., Joutsensaari, J., Koop, T., Kannosto, J., Yli-Pirila, P., Leskinen, J.,
from terpene ozonolysis. 2. Effect of NOx concentration. Environ. Sci. Technol. 39, Makela, J.M., Holopainen, J.K., Poschl, U., Kulmala, M., Worsnop, D.R.,
7046–7054. Laaksonen, A., 2010. An amorphous solid state of biogenic secondary organic aerosol
Pye, H.O., Pinder, R.W., Piletic, I.R., Xie, Y., Capps, S.L., Lin, Y.-H., Surratt, J.D., particles. Nature 467, 824–827.
Zhang, Z., Gold, A., Luecken, D.J., 2013. Epoxide pathways improve model von Hessberg, C., von Hessberg, P., Poschl, U., Bilde, M., Nielsen, O.J., Moortgat, G.K.,
predictions of isoprene markers and reveal key role of acidity in aerosol formation. 2009. Temperature and humidity dependence of secondary organic aerosol yield
Environ. Sci. Technol. 47, 11056–11064. from the ozonolysis of β-pinene. Atmos. Chem. Phys. 9, 3583–3599.
Pye, H.O.T., Luecken, D.J., Xu, L., Boyd, C.M., Ng, N.L., Baker, K.R., Ayres, B.R., Bash, J. Wildt, J., Mentel, T.F., Kiendler-Scharr, A., Hoffmann, T., Andres, S., Ehn, M., Kleist, E.,
O., Baumann, K., Carter, W.P.L., Edgerton, E., Fry, J.L., Hutzell, W.T., Schwede, D.B., Müsgen, P., Rohrer, F., Rudich, Y., Springer, M., Tillmann, R., Wahner, A., 2014.
Shepson, P.B., 2015. Modeling the current and future roles of particulate organic Suppression of new particle formation from monoterpene oxidation by NOx. Atmos.
nitrates in the southeastern United States. Environ. Sci. Technol. 49, 14195–14203. Chem. Phys. 14, 2789–2804.
Reis, S., Pinder, R.W., Zhang, M., Lijie, G., Sutton, M.A., 2009. Reactive nitrogen in Ye, P., Ding, X., Hakala, J., Hofbauer, V., Robinson, E.S., Donahue, N.M., 2016. Vapor
atmospheric emission inventories. Atmos. Chem. Phys. 9, 7657–7677. wall loss of semi-volatile organic compounds in a Teflon chamber. Aerosol Sci.
Renbaum, L.H., Smith, G.D., 2009. Organic nitrate formation in the radical-initiated Technol. 50, 822–834.
oxidation of model aerosol particles in the presence of NOx. Phys. Chem. Chem. Yu, K.Y., Zhu, Q., Du, K., Huang, X.F., 2019. Characterization of nighttime formation of
Phys. 11, 8040–8047. particulate organic nitrates based on high-resolution aerosol mass spectrometry in an
Rollins, A.W., Browne, E.C., Min, K.E., Pusede, S.E., Wooldridge, P.J., Gentner, D.R., urban atmosphere in China. Atmos. Chem. Phys. 19, 5235–5249.
Goldstein, A.H., Liu, S., Day, D.A., Russell, L.M., Cohen, R.C., 2012. Evidence for Zaveri, R.A., Shilling, J.E., Zelenyuk, A., Liu, J.M., Bell, D.M., D’Ambro, E.L., Gaston, C.,
NOx control over nighttime SOA formation. Science 337, 1210–1212. Thornton, J.A., Laskin, A., Lin, P., Wilson, J., Easter, R.C., Wang, J., Bertram, A.K.,
Saathoff, H., Naumann, K.H., Mohler, O., Jonsson, Å.M., Hallquist, M., Kiendler- Martin, S.T., Seinfeld, J.H., Worsnop, D.R., 2018. Growth kinetics and size
Scharr, A., Mentel, T.F., Tillmann, R., Schurath, U., 2009. Temperature dependence distribution dynamics of viscous secondary organic aerosol. Environ. Sci. Technol.
of yields of secondary organic aerosols from the ozonolysis of α-pinene and 52, 1191–1199.
limonene. Atmos. Chem. Phys. 9, 1551–1577. Zhang, Q., Jimenez, J.L., Canagaratna, M., Allan, J., Coe, H., Ulbrich, I., Alfarra, M.,
Sakulyanontvittaya, T., Guenther, A., Helmig, D., Milford, J., Wiedinmyer, C., 2008. Takami, A., Middlebrook, A., Sun, Y., 2007. Ubiquity and dominance of oxygenated
Secondary organic aerosol from sesquiterpene and monoterpene emissions in the species in organic aerosols in anthropogenically-influenced Northern Hemisphere
United States. Environ. Sci. Technol. 42, 8784–8790. midlatitudes. Geophys. Res. Lett. 34, L13801.
Sander, S., Abbatt, J., Barker, J., Burkholder, J., Friedl, R., Golden, D., Huie, R., Kolb, C., Zhang, X., Schwantes, R.H., McVay, R.C., Lignell, H., Coggon, M.M., Flagan, R.C.,
Kurylo, M., Moortgat, G., 2011. Chemical kinetics and photochemical data for use in Seinfeld, J.H., 2015. Vapor wall deposition in Teflon chambers. Atmos. Chem. Phys.
atmospheric studies, evaluation number 17. JPL Publ. 10, 1–11. 15, 4197–4214.
Shrivastava, M., Cappa, C.D., Fan, J.W., Goldstein, A.H., Guenther, A.B., Jimenez, J.L., Zhang, Y., Tang, L., Sun, Y., Favez, O., Canonaco, F., Albinet, A., Couvidat, F., Liu, D.,
Kuang, C., Laskin, A., Martin, S.T., Ng, N.L., Petaja, T., Pierce, J.R., Rasch, P.J., Jayne, J.T., Wang, Z., 2017. Limited formation of isoprene epoxydiols-derived
Roldin, P., Seinfeld, J.H., Shilling, J., Smith, J.N., Thornton, J.A., Volkamer, R., secondary organic aerosol under NOx-rich environments in Eastern China. Geophys.
Wang, J., Worsnop, D.R., Zaveri, R.A., Zelenyuk, A., Zhang, Q., 2017. Recent Res. Lett. 44, 2035–2043.
advances in understanding secondary organic aerosol: implications for global Zhao, D.F., Kaminski, M., Schlag, P., Fuchs, H., Acir, I.H., Bohn, B., H€aseler, R., Kiendler-
climate forcing. Rev. Geophys. 55, 509–559. Scharr, A., Rohrer, F., Tillmann, R., Wang, M.J., Wegener, R., Wildt, J., Wahner, A.,
Slade, J.H., de Perre, C., Lee, L., Shepson, P.B., 2017. Nitrate radical oxidation of Mentel, T.F., 2015. Secondary organic aerosol formation from hydroxyl radical
γ-terpinene: hydroxy nitrate, total organic nitrate, and secondary organic aerosol oxidation and ozonolysis of monoterpenes. Atmos. Chem. Phys. 15, 991–1012.
yields. Atmos. Chem. Phys. 17, 8635–8650. Zheng, Y., Unger, N., Hodzic, A., Emmons, L., Knote, C., Tilmes, S., Lamarque, J.F.,
Yu, P., 2015. Limited effect of anthropogenic nitrogen oxides on secondary organic
aerosol formation. Atmos. Chem. Phys. 15, 13487–13506.

11

You might also like