PreprintPDF Available

A comprehensive evaluation of enhanced temperature influence on gas and aerosol chemistry in the lamp-enclosed oxidation flow reactor (OFR) system

Authors:
  • Chinese academy of sciences,Geochemistry institue at Guangzhou

Abstract

Oxidation flow reactors (OFRs) have been widely used to investigate the formation of secondary organic aerosol (SOA). However, the UV lamps that are commonly used to initiate photochemistry in OFRs can lead to increases in the reactor temperature with consequences that have not been assessed in detail. In this study, we systematically investigated the temperature distribution inside an Aerodyne Potential Aerosol Mass OFR and the associated impacts on flow and chemistry arising from lamp heating. A lamp-induced temperature enhancement was observed, which was a function of lamp driving voltage, number of lamps, lamp types, OFR residence time, and positions inside OFR. Under common OFR operational conditions (e.g., < 5 days of equivalent atmospheric OH exposure under low-NOx conditions), the temperature enhancement was usually within 1–5 °C. Under extreme (but less commonly used) settings, the heating could reach 15 °C. The influence of the increased temperature over ambient conditions on the flow distribution, gas, and condensed phase chemistry inside OFR was evaluated. We found that the increase in temperature changes the flow field, leading to a reduced tail on the residence time distribution and corresponding oxidant exposure due to faster recirculation. According to simulation results from a box model using radical chemistry, the variation of absolute oxidant concentration inside of OFR due to temperature increase was small (<5 %). The temperature influences on existing and newly formed OA were also investigated, suggesting that the increase in temperature can impact the yield, size, and oxidation levels of representative biogenic and anthropogenic SOA types. Recommendations for temperature-dependent SOA yield corrections and OFR operating protocols that mitigate lamp-induced temperature enhancement and fluctuations are presented. We recommend blowing air around the outside of the reactor with fans during OFR experiments to minimize the temperature increase inside OFR. Temperature increases are substantially lower for OFRs using less powerful lamps than the Aerodyne version.
1
A comprehensive evaluation of enhanced temperature influence on
gas and aerosol chemistry in the lamp-enclosed oxidation flow
reactor (OFR) system
Tianle Pan1,2,3,4,5, Andrew T. Lambe6, Weiwei Hu1,2,4,5, Yicong He7,a, Minghao Hu8, Huaishan
Zhou1,2,3,4,5, Xinming Wang1,2,4,5, Qingqing Hu9, Hui Chen9, Yue Zhao10, Yuanlong Huang11, Doug R.
5
Worsnop6,12, Zhe Peng13,14, Melissa A. Morris13,14, Douglas A. Day13,14, Pedro Campuzano-Jost13,14,
Jose-Luis Jimenez13,14, Shantanu H. Jathar6
1State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences,
Guangzhou 510640, China
2CAS Center for Excellence in Deep Earth Science, Guangzhou, 510640, China
10
3Chinese Academy of Sciences University, Beijing 100049, China
4Guangdong-Hong Kong-Macao, Joint Laboratory for Environmental Pollution and Control, Guangzhou Institute of
Geochemistry, Chinese Academy of Science, Guangzhou 510640, China
5Guangdong Provincial Key Laboratory of Environmental Protection and Resources Utilization, Chinese Academy of
Science, Guangzhou 510640, China
15
6Aerodyne Research Inc., Billerica, Massachusetts, 01821, United States
7Department of Mechanical Engineering, Colorado State University, Fort Collins, Colorado 80523, United States
8China-UK Low Carbon College, Shanghai Jiao Tong University, Shanghai, 201306, China
9Key Laboratory of Organic Compound Pollution Control Engineering, School of Environmental and Chemical
Engineering, Shanghai University, 200444, Shanghai, China
20
10School of Environmental Science and Engineering, Shanghai Jiao Tong University, Shanghai, 200240, China
11Department of Environmental Science and Engineering, California Institute of Technology, Pasadena, California 91125,
United States
12Institute for Atmospheric and Earth System Research (INAR) / Physics, Faculty of Science, University of Helsinki,
Helsinki, 00014, Finland
25
13Cooperative Institute for Research in the Environmental Sciences (CIRES), University of Colorado at Boulder, Boulder,
Colorado, 80309, United States
14Department of Chemistry, University of Colorado at Boulder, Boulder, Colorado 80309, United States
anow at: State Key Joint Laboratory of Environmental Simulation and Pollution Control, School of
Environment, Tsinghua University, Beijing 100084, China
30
Correspondence to: Weiwei Hu (weiweihu@gig.ac.cn)
Abstract. Oxidation flow reactors (OFRs) have been widely used to investigate the formation of secondary organic aerosol
(SOA). However, the UV lamps that are commonly used to initiate photochemistry in OFRs can lead to increases in the
reactor temperature with consequences that have not been assessed in detail. In this study, we systematically investigated
the temperature distribution inside an Aerodyne Potential Aerosol Mass OFR and the associated impacts on flow and
35
chemistry arising from lamp heating. A lamp-induced temperature enhancement was observed, which was a function of
lamp driving voltage, number of lamps, lamp types, OFR residence time, and positions inside OFR. Under common OFR
operational conditions (e.g., < 5 days of equivalent atmospheric OH exposure under low-NOx conditions), the temperature
enhancement was usually within 1-5 °C . Under extreme (but less commonly used) settings, the heating could reach 15 °C .
The influence of the increased temperature over ambient conditions on the flow distribution, gas, and condensed phase
40
chemistry inside OFR was evaluated. We found that the increase in temperature changes the flow field, leading to a reduced
tail on the residence time distribution and corresponding oxidant exposure due to faster recirculation. According to
simulation results from a box model using radical chemistry, the variation of absolute oxidant concentration inside of OFR
due to temperature increase was small (<5%). The temperature influences on existing and newly formed OA were also
investigated, suggesting that the increase in temperature can impact the yield, size, and oxidation levels of representative
45
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
2
biogenic and anthropogenic SOA types. Recommendations for temperature-dependent SOA yield corrections and OFR
operating protocols that mitigate lamp-induced temperature enhancement and fluctuations are presented. We recommend
blowing air around the outside of the reactor with fans during OFR experiments to minimize the temperature increase inside
OFR. Temperature increases are substantially lower for OFRs using less powerful lamps than the Aerodyne version.
1 Introduction
50
Secondary organic aerosol (SOA) can account for 60-95% of OA and 10-75% of submicron particles (Jimenez et al.,
2009), thus, strongly impacting air quality (Huang et al., 2014), climate (Myhre, 2013; Poschl, 2005) and human health
(Nel, 2005; Feng et al., 2016; Nault et al., 2021). Elucidating the formation mechanism of SOA is crucial for clarifying its
environmental impact (Ziemann and Atkinson, 2012; Hallquist et al., 2009; Klyta and Czaplicka, 2020). To investigate
SOA formation, chambers (Hildebrandt et al., 2009; Cocker et al., 2001; Paulsen et al., 2005; Wang et al., 2014; Carter et
55
al., 2005; Martin-Reviejo and Wirtz, 2005; Rollins et al., 2009; White et al., 2018; Zong et al., 2023) and flow tubes
(Cooper and Abbatt, 1996; George et al., 2007; Hanson and Lovejoy, 1995; Robbins and Cadle, 2002; Katrib et al., 2005;
Knopf et al., 2005; Ezell et al., 2010) have been commonly used in the laboratory or investigating secondary formation in
the early periods of past several decades. In the last decade, due to the fast development of online measurement techniques,
different types of oxidation flow reactors (OFR), which are portable and can be used in field studies to explore SOA
60
formation under ambient conditions, have been developed and widely deployed (Kang et al., 2007; Watne et al., 2018;
George et al., 2007; Peng and Jimenez, 2020; Lambe et al., 2011b; Shah et al., 2020; Saha et al., 2018; Xu and Collins,
2021; Chu et al., 2016; Simonen et al., 2017; Li et al., 2019; Keller et al., 2022 and references therein).
The potential aerosol mass (PAM) reactors, initially distributed by Prof. Bill Brune (the Penn State PAM) and later
modified and commercialized by Aerodyne Research Inc. are the most widely used OFRs for studying SOA formation and
65
evolution under ambient environments (Peng and Jimenez, 2020). The PAM reactor was first proposed and designed by
Kang et al. (Kang et al., 2007) and has been successfully deployed in a variety of field studies including forests (Hu et al.,
2016; Palm et al., 2016; Palm et al., 2017; Palm et al., 2018; Sumlin et al., 2021), urban areas (Ortega et al., 2016; Liu et
al., 2018; Chen et al., 2021; Hu et al., 2022; Sbai et al., 2021; Park and Kim, 2023; Xu et al., 2022; Park et al., 2019), rural
sites (Ahlberg et al., 2019; Hodshire et al., 2018; Hu et al., 2016), tunnel/roadside studies (Liu et al., 2019; Liao et al., 2021;
70
Tkacik et al., 2014; Saha et al., 2018), as well as in multiple laboratory studies (Kang et al., 2007; Lambe et al., 2011b;
Bahreini et al., 2012; Bruns et al., 2015; Sengupta et al., 2018; Kramer et al., 2019; Cheng et al., 2021; Lei et al., 2022;
Srivastava et al., 2023 and references therein). In the PAM, high concentrations of the OH radicals, which is the major
oxidant for SOA formation in ambient air, can be generated quickly by initiating the photochemistry of O3 and H2O with
UV lamps (185/254 nm) mounted inside of a flow tube (Lambe et al., 2011b; Li et al., 2015; Peng et al., 2015).
75
Earlier studies using the Penn State PAM reported a temperature increase of 2 °C . However, several recent studies
using the PAM show that the UV lamps can increase the internal temperature of that reactor by 0-15 °C (up values are
under extremely high voltage settings) above ambient temperature (Lambe et al., 2011a; Chen et al., 2013; Tkacik et al.,
2014; Lambe et al., 2019; Charnawskas et al., 2017). However, a full investigation of the lamp heating effect on the
temperature distribution inside of PAM-OFR and clarification of its influences on the flow distribution inside of PAM-
80
OFR has not been reported yet. The temperature is a key parameter for gas diffusion and gas/aerosol partitioning (Pankow,
1994; Donahue et al., 2006). The increase in the temperature inside an OFR can result in the formation or deepening of
recirculating flow, which leads to a shorter residence time and a broader residence time distribution (Huang et al., 2017;
Lambe et al., 2019). Numerous chamber studies have found that higher temperatures may substantially decrease the SOA
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
3
yield (Gao et al., 2022; Lamkaddam et al., 2017; Boyd et al., 2017; Price et al., 2016; Clark et al., 2016; Tillmann et al.,
85
2010; Pathak et al., 2007) and influence SOA chemical composition (Jensen et al., 2021; Gao et al., 2022; Simon et al.,
2020; Li et al., 2020a; Kristensen et al., 2020; Quéver et al., 2019; Denjean et al., 2015). Thus, the increase of temperature
inside of PAM-OFR, which could lead to extra uncertainty for SOA simulation and study, should be further assessed and
clarified.
In this study, we systemically investigate the effects of lamp-induced heating on PAM-OFR measurements. Based on
90
computational fluid dynamics (CFD) simulation, we show how the temperature affects the flow and average OH exposure
inside the PAM-OFR. Two box models were applied to illustrate how the enhanced temperature impacts the chemistry of
gas-phase reactions and SOA formation inside PAM-OFR. The model results can be used as a rough reference for
correcting the related experimental results due to enhanced temperature. Finally, we propose the use of fans to cool the
OFR and an online sampling strategy that alternates high and low driving voltages at varied lamp settings to reduce the
95
uncertainty caused by the varied heating effect.
2 Methods
2.1 Oxidation flow reactor (OFR)
The PAM OFR (Aerodyne Research, Inc.) used in this study is a 13 L horizontal aluminum cylindrical chamber (46
cm long × 22 cm ID). Four low-pressure germicidal mercury (Hg) lamps are isolated from the sample flow by type 214
100
quartz tubes (Fig. S1). The Hg UV fluorescent lamps used in this study are commercial products (GPH436T5VH/4 or
GPH436T5L/4, Light Sources, Inc.) which are the default light bulbs chosen by ARI for the OFR (Fig. S2). In addition,
the temperature within an OFR with the Penn State low-pressure Hg UV lamps (model no. 82-9304-03, BHK Inc.), which
are also widely used in PAM-OFR (Khalaj et al., 2021; Xu and Collins, 2021; Lambe et al., 2011b; Siemens et al., 2022;
Hu et al., 2016; Ortega et al., 2013; Link et al., 2016; Palm et al., 2016; Kang et al., 2018; Mitroo et al., 2018; Kang et al.,
105
2007), are also measured. For the Light Sources lamps, the light intensity of each lamp was changed through the AC voltage
input to the lamp ballast, controlled by a computer with settings ranging from 0 V (minimum) to 10 V (maximum, full AC
output). For BHK lamps, the maximum setting voltage is 5 V (full AC output). At maximum voltage, the best estimation
of supply power is 8.9 W (see sect. S1 for details) to each lamp for the Light Source lamp and 6.3 W for the BHK lamp.
When the lamps are turned on, a pure nitrogen purge gas flows through the lamp sleeve at a flow rate of 0.2-0.3 L min-1.
110
The lamp sleeve is defined as the space between the lamps and the wall of the quartz tube. This is done to avoid the
accumulation of oxidation products on the surface of the lamps and to somewhat cool the UV lamps (minor effect, as
discussed below). On the inner surface of the front plate of the OFR, a hexagon nut is connected to the center inlet. The
side of the nut is drilled with holes to promote axial mixing of the sample flow, and is hereafter referred to as an inlet
diffuser (Fig. S1c) (Mitroo et al., 2018). Two mesh screens, used to block debris and insects and make the air flow more
115
even (helping to break up eddies from outside), are installed inside of the front and back plates, respectively (Fig. S1b). A
detailed schematic plot of the OFR used in this study can be found in Fig. 1.
2.2 Measurement of OFR temperature
The temperature distribution inside OFR was measured with multiple temperature sensors. The types and models of
temperature sensors used are listed in Table S1. To assess the accuracy of temperatures measured by different temperature
120
sensors, a comparison experiment was carried out under different ambient environments. We found good agreement (within
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
4
2 °C ) among different temperature sensors, supporting the robustness of the temperatures measured in this study (Fig. S3).
Detailed information on experimental settings for temperature measurements inside the OFR is summarized in Fig. 1 and
Table S2. Briefly, we measured the air temperature inside OFR at varied positions (vertically and horizontally) under
different lamp settings (e.g., number of lamps, lamp types, and lamp intensity) and flow rates. To achieve this, an external
125
temperature sensor was placed directly inside the OFR and was moved manually from outside. The values reported in this
study were obtained after the temperature inside OFR was stable for ~20 min after switching OFR settings. During the
measurement, the inlet diffuser and the mesh screen in the front plate of OFR or in the backplate needed to be removed to
extend the thermocouple inside the OFR. We have tested if the removal of the mesh screen would change the temperature
field inside OFR and found a negligible impact (Fig. S4 (removed the back plate mesh screen) vs. Fig. 5 (removed the inlet
130
diffuser and front plate mesh screen)). We also covered the sensor with shielding paper to avoid the potential influences of
optical radiation from the UV lamps on temperature detection during this process. In addition to the inside temperature of
OFR, the temperature of the inner/external surface of OFR, and the temperature in the gas output of OFR were also
measured for comparison. A sensor for detecting temperature and relative humidity (RH) (referred to as a “primary T/RH
sensor”, Sensirion SHT21, Sensirion AG) was mounted on the back plate at the factory. Thus, the default primary T/RH
135
sensor was used for temperature measurement for all the experiments.
Figure 1: Schematic plot for temperature measurement in the oxidation flow reactor of this study. The center inlet,
nut, and mesh screen near the front plate were removed when the temperature sensor was probed in the front
140
direction. The information of different temperature sensors used can be found in Table S1.
2.3 Model simulations on temperature distribution and flow field inside OFR
Computational Fluid Dynamics (CFD) simulations were performed using the ANSYS Fluent software (Version 14.5)
in three dimensions to simulate the temperature distribution and flow field in OFR. ANSYS has been used to simulate the
flow field inside flow reactors in past studies (Li et al., 2019; Ihalainen et al., 2019). The temperature distribution and flow
145
field under unheated and heated conditions were both simulated with the CFD model. The simulated experimental condition
is 5 L min-1 of air as the carrier gas and 0.3 L min-1 of SO2 which was injected for 2 s as the tracer gas. The convergence
of this model was defined when the residuals of physical quantities (e.g., pressure, temperature, velocity, density, and
viscosity, etc.) were below 0.001.
For the unheated condition, the more advanced “realizable k-epsilon turbulence model” was used (Shih et al., 1995).
150
The simulation was solved by using the pressure-based SIMPLEC algorithm, which is a commonly used method in CFD
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
5
models for simulating incompressible flow problems (Patankar and Spalding, 1972), as OFR conditions with lights were
off. For the heated condition, the thermal boundary was set to a series of fixed temperatures, based on those directly
measured in the OFR. When the driving voltage of the OFR lamps was all set to be 5 V, the model settings applied were
55 °C for the surface of the quartz UV lamp sleeves, 35 °C for the inner wall surface, and 26 °C for the carrier gas. When
155
all the voltages of the OFR lamps were 10 V, those values were set to 62 °C , 40.5 °C , and 23 °C , respectively. The lower
carrier gas temperature under 10 V conditions vs, 5V was caused by the different room temperatures in each real experiment.
The viscosity and thermal conductivity of mixing gas were calculated using the mass-weighted-mixing-law (Ni et al., 2010).
Least Squares Cell-Based method and species transport model were used for the solution (Ghia et al., 1982).
2.4 KinSim kinetic model for gas-phase reactions
160
The influence of temperature on gas-phase reactions was modeled with a box model (KinSim 4.14 in Igor Pro. 6.37)
with the OFR radical mechanisms used in Li et al. (2015), Peng et al. (2015) and Peng et al. (2019). All the gas-phase
reactions were obtained from the JPL chemical kinetic data evaluation, which includes temperature as a variable for reaction
rate calculation (Burkholder et al., 2019). In the model, we assumed a mixing volume of 800 ppb SO2 injection into OFR
under a constant 2.2% water vapor mixing ratio (RH=70% at 25 °C ) based on calibration experiment settings. This
165
corresponds to an OH reactivity (OHR) of the incoming air of 20 s-1. The simulated SO2 output concentration in the model
was weighted by the measured residence time distribution (RTD) at a flow rate of 5 L min-1 (~700 s). Good agreement
between modeled SO2 decay based on the KinSim model and measured SO2 results has been shown in Hu et al. (2022).
The mechanism of OFR185 mode, in which O3 and OH radicals were generated by the photolysis of O2 and H2O inside
OFR, was applied in the model. Note that the ratio of photon flux between 185 and 254 nm from lamps produced by Light
170
Source Inc. does not change as a function of their intensity (Rowe et al., 2020), while these ratios do vary with intensity
for the BHK lamps (Li et al., 2015). In the model, the photon flux ratio of 185/254 nm was set to be constant (5%). Results
for 25 to 40 °C (binned with 5 °C ) were modeled, which covers the typical temperature range inside OFR under most
lamp setting cases. These simulations were done with RTD obtained under unheated conditions in OFR (25 °C ). Since the
residence time distribution (RTD) inside OFR can also be influenced by temperature (Lambe et al., 2019), we also show
175
the model results with measured RTD at 40 °C .
2.5 SOM model for SOA formation
The statistical oxidation model (SOM) is a kinetic, process-level model to simulate SOA formation by accounting for
gas-particle partitioning, multi-generational oxidation (fragmentation and functionalization), and autoxidation for highly
oxygenated molecules (HOMs) (Cappa and Wilson, 2012; Eluri et al., 2018; He et al., 2021; He et al., 2022). The SOM
180
model has been shown to successfully capture the evolution of SOA formation and oxidation in multiple laboratory and
field studies (Cappa and Wilson, 2012; Jathar et al., 2015; Eluri et al., 2018; Akherati et al., 2020; He et al., 2021; He et
al., 2022). In the SOM model, the oxidation of a VOC precursor is tracked by its evolution within a carbon−oxygen grid.
The volatilities and reactivity (kOH) of organic species in each grid are calculated based on their carbon (NC) and oxygen
numbers (NO) (Cappa and Wilson, 2012; Jathar et al., 2015; Eluri et al., 2018). In total, six adjustable parameters are used
185
to determine the reaction probabilities and volatility of the grid species. Four parameters (p1-p4, Table S3) indicate the
number of oxygens added per functionalization reaction, and two other parameters describe the probability of fragmentation
(, Table S3) and the decrease in vapor pressure (, Table S3) per oxygen atom added to the carbon backbone
(Jathar et al., 2016; Akherati et al., 2019).
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
6
In this study, we used the SOM model coupled with the TwO-Moment Aerosol Sectional model (TOMAS) (Adams
190
and Seinfeld, 2002; Pierce et al., 2007) to simulate the temperature influences on the yield, oxidation state, and size
distributions of newly formed SOA in the OFR. In the model, the temperature dependence is mainly calculated based on
the ClausiusClapeyron Equation:
󰇛

󰇡

󰇢󰇜 (1)
where (8.314 J mol-1 K-1) is the ideal gas constant, and
(Pa) are the saturation pressure of species i at target
195
temperature and reference temperature (298 K used here), respectively. The
can be calculated based on the following
equation (2):

 (2)
where  (g mole−1) is the molecular weight of organic species i and
(μg m-3) is the saturation
concentration at 298 K. The  and
are calculated based on the number of C () and O () and  of
200
species i (Eluri et al., 2018). The
(kJ mol-1) is the evaporation enthalpy of species i calculated as followed equation
(Epstein et al., 2010):
 
 (3)
In addition, constant
of 80, 100, 120 and 150 kJ mol-1, derived from analysis of field data, was also tested in
the model (Saha et al., 2017; Cappa and Jimenez, 2010; Louvaris et al., 2017).
205
In this study, SOA formation from four typical VOC precursors including dodecane, α-pinene, toluene, and m-xylene
under different OA concentrations (1-80 μg m-3) and NOx conditions (low NOx vs high NOx) was modeled. Five stages
of temperature ranging from 20-40 °C were simulated. Similar to the gas-phase simulations, results were also calculated
based on the measured RTD obtained at both ~25 and ~40 (for the modeled yields under 40 ) with a total model
time of 700 s. Additionally, we also considered the highly oxygenated organic molecules (HOMs) in this model that
210
contribute to new particle formation (He et al., 2021; Bianchi et al., 2019). The HOM yields and other detailed information
for input parameters can be found in Table S3.
3 Results and discussions
3.1 Heating effect inside OFR
To characterize the temperature distribution in OFR, we measured the temperature as a function of lamp driving voltage,
215
number of lamps used, flow rate, lamp types, and different space positions.
3.1.1 Enhanced temperature vs lamp driving voltage (OH exposures) and flow rate
In general, we observed a systematic increase in temperature inside the OFR with Light Source lamps. As shown in
Fig. 2a, the enhanced temperature in the OFR increases as a function of lamp driving voltage and, hence, OH exposures.
When the driving voltage on the 2-lamp setup was changed in increasing fashion (from 0 V to 3 V, Light Source lamp),
220
the temperature enhancement in the OFR was generally lower than 5 °C. 3 V is roughly equivalent to a photochemical
aging time of around 5 days (assuming the water mixing ratio is 1.88%, RH≈60%, external OH reactivity=30 s-1). For a
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
7
higher voltage range (from 3 V to 10 V, which is equivalent to a photochemical aging time from 5 days to two weeks), the
temperature inside the OFR increased by 10 °C . Note that the relationship between the lamp driving voltage and OH
exposure varied with the number and type of UV lamps used, as well as how the OFR was operated, as shown in Fig. S5.
225
When four lamps were turned on and the driving lamp voltage was increased from 2 to 10 V, the temperature increased by
5-20 °C inside the OFR for the laboratory tests (Fig. 2a and 3a). These results support that the temperature increase inside
of the OFR is mainly due to the heat from the lamps and was further supported by Fig. 3b, which shows that the temperature
enhancement inside the OFR decreased as a function of the flow rate from 3 L min-1 to 7 L min-1 (Fig. 3b). The anti-
correlation between the temperature enhancement and the flow rate was mainly due to the higher heat capacity provided
230
by the increased air flow, thus needing a lower temperature increase to carry the excess heat from the lamps.
Figure 2: (a) Enhanced temperature inside OFR (Measured temperature in OFR minus ambient temperature) as a
function of the lamp driving voltage using two lamps and four lamps in the laboratory studies. Note that without
specific notification, all the temperature measurement results shown in the main text are obtained with lamps
235
produced by Light Source Inc. The relationship between lamp driving voltage and OH exposures for using two
lamps and four lamps can be found in Fig. S5. The OH exposures were calculated based on the empirical parameters
in Hu et al. (2022) assuming a mixing water ratio of around 1.88% and external OH reactivity of 30 s-1. The
equivalent aging days (top axis) were estimated by assuming ambient OH concentrations are around 1.5x106
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
8
molecules cm-3. The lamp type applied here is uncovered for field studies with four lamps in Fig. 2b and 80% covered
240
for other conditions. The measured position of temperature is in the centerline with 300 nm probing. The flow rate
is 5 L min-1. (b) Enhanced temperature inside OFR as a function of light settings in the field studies. Two lamps
were applied in the low NOx conditions, while the four lamps were for high NOx conditions. The high NOx reaction
regimes were achieved by inputting extra N2O as illustrated in Peng et al. (2018) and Lambe et al. (2017).
245
Figure 3: (a) The measured temperature under different lamp driving voltage and (b) at different flow rates for
Light Source lamp. The temperature was measured in the centerline at different probing positions. Black lines and
gray shadows represent the temperature measured by the primary T/RH sensor. For panel (a) four lamps were all
turned on during measurement. The flow rate is 5 L min-1; For panel (b), four lamps were set in 10 V.
When the lamp driving voltage was set in a monotonically decreasing fashion (e.g., from 10 V to 0 V), the maximum
250
temperature increase did not occur at the highest voltage setting of lamps (10 V), but in the range of 4-6 V, as shown in
Fig. 2a and b. This inconsistency is mainly due to the lamps starting at 10 V with colder conditions (e.g., room temperatures
or lower voltage settings), meanwhile, the OFR reactor has a thermal mass that needs time to accumulate or dissipate heat.
Fig. 4 shows that the entire system needs more than 100 minutes to achieve a thermal steady state. However, a much shorter
time (15-24 minutes) is usually applied in the laboratory and field studies when voltages are systematically varied (Link et
255
al., 2016; Murschell and Farmer, 2018; Hu et al., 2016; Ortega et al., 2016; Palm et al., 2016; Palm et al., 2018; Saha et al.,
2018; Shah et al., 2020; Hu et al., 2022). These short constant lamp power times result in reduced temperature variation
between different power settings, as well as inconsistency between maximum temperature peak with maximum power
setting. To be clear, the time series of measured temperature when the lamp driving voltage was set under a representative
lamp voltage cycle (decreasing fashion: 10-6-5-4-2-0 V) are displayed in Fig. 5. Two scenarios with two and four lamps
260
being turned on were both shown. The reason that the lamp voltage starts with 10 V in this typical setting is to maximize
the efficiency of filament ignition in UV fluorescent lamps. Even though the thermal steady state was not achieved at each
voltage setting, temperature increases of 2-8 °C for two lamps and 6-16 °C for four lamps were observed. In general, the
temperature enhancement inside OFR is a function of the voltage profile and the time spent at each voltage. Maximum
SOA formation is typically observed after 1-2 days of equivalent age (Ortega et al., 2016; Palm et al., 2016; Hu et al.,
265
2022). In most cases, the temperature enhancement inside OFR is generally less than 5 °C below 5-7 days of equivalent
atmospheric OH exposure under laboratory and field studies as shown in Fig. 2a and 2b. The main exceptions involve
specific OFR modes that incorporate N2O or isopropyl nitrite precursors to establish high-NOx conditions (Lambe et al.,
2017; Lambe et al., 2019). In these cases, four lamps are beneficial to compensate for the absorption of 185 nm radiation
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
9
by N2O (Peng et al., 2018) that hinders HOx production, or the relatively weak absorption of UVA radiation by isopropyl
270
nitrite, respectively. This also makes the radiation within the OFR more uniform, as when only two lamps are used, the
area with another two lamps off would be darker due to the absorption of radiation.
Figure 4: The temperature variation under 10 V driving voltage for 4 lamps as a function of time. The start time is
the lamps being turned on. The temperature sensors were set at a depth of 300 mm in the central line to measure
275
the temperature inside OFR. The flow rate is 5 L min-1. The temperature of the external surface of OFR is also
measured at the top and bottom positions. The temperature measured by the default OFR sensor installed in the
backplate is also shown in the black line. The purple dotted lines represent the ambient temperature.
3.1.2 Enhanced temperature vs lamp types.
The temperature enhancement as a function of different types of lamps was also investigated here. The four types of
280
tested lamps are shown in Fig. S2. Fig. S6 shows similar temperature distribution inside OFR no matter whether 80%
covered or uncovered lamps, as well as no matter whether only 254 nm or 254/185 nm lamps were used. These results
suggest that the heating inside OFR is mainly from the heat transfer of the hot quartz sleeve (heated by the lamps) but not
from their optical radiation. Thus, different types of lamps impact mostly the OH exposures, but not the heating effect. The
three types of lamps applied here are all from the same manufacturer (Light Source Inc.) and, therefore have a similar value
285
of power supplied to each lamp (8.9 W). In addition, we also tested the temperature enhancement in the OFR with lamps
from another producer (BHK Inc.). An enhanced temperature of 6 °C inside the OFR was found at the centerline when two
BHK lamps at full power (Fig. S7b) were used in an OFR with a flow rate of 4.5 L min-1. The less enhanced temperature
with BHK lamps is probably due to the lower input heating energy (6.3 W per lamp) than the lamp from Light Sources.
For the lower driving voltage of 0.95 V for BHK lamps as run during a field campaign, a temperature enhancement of 1-
290
2 °C with an equivalent aging time of around 1.5 days was observed (Fig. S7a and S7b). For that campaign, the OFR was
operated mostly continuously at ~1.5 days equivalent aging, since most SOA formation often is observed at these moderate
exposures, and is low enough that heterogeneous oxidation is not yet significant (e.g., Palm et al. (2016), Ortega et al.
(2016) and Hu et al. (2016)).
3.1.3 Mapping out the temperature enhancement inside the OFR
295
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
10
In this section, the enhanced temperatures inside of the OFR as a function of position are explored. In our case, the
voltages of four lamps are set to be 5 V (~30 days' equiv. aging time and flow rate of 5 L min-1) the temperature in the OFR
is generally 9-15 °C higher (Fig. 6a) than ambient (26 °C ). In the vertical axial direction, the enhanced temperature is
higher at the top position (~14 °C ) than at the bottom (~9 °C ) due to the warm air ascending in OFR resulting from its
lighter density. We also found a similar conclusion for BHK lamps (Fig. S7) with lower temperature enhancement and
300
smaller temperature difference in the vertical direction for the same equivalent aging. Horizontally, the temperature
distribution is symmetrical, which shows a slightly lower temperature value in the middle and enhanced temperature on
the edges (Fig. 6b). These temperature values were measured until their reading varied no more than 0.1 °C . The lower
center temperature is probably caused by the faster flow rate (shorter residence time) of air mass in the middle due to the
removal of the inlet diffuser and the longer distance from the lamps. However, the largest temperature gradient at different
305
horizontal axial positions is within 2-3 °C , suggesting the general temperature distribution measured here is similar to the
distributions when the inlet diffuser is installed. For different probing depths, Fig. 3 generally shows the enhanced
temperature is lower near the inlet and higher from the middle position to the back in the central line, which is mainly due
to that the airflow flowing from the inlet to the outlet is being warmed, as well as the set position of lamps has a gap with
the front plate with the input inlet, as demonstrated in Fig. 1.
310
Figure 5: Temperature variation under different lamp driving voltage cycled as 10-6-5-4-2-0 V for: (a) Two
symmetrical lamps on the diagonal direction were used and the OFR backplate sensor was extended into OFR (from
the backplate) at a depth of 300 mm from the inlet. (b) Four lamps were used and the OFR sensor was set in the
backplate. Each lamp setting lasts for 16 min. The flow rate of sampling air is 5 L min-1 (residence time is 167 s).
315
3.1.4 Loss of heating energy in OFR
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
11
The gain of the heating energy inside the OFR generally comes from the UV lamps, while the energy loss inside the
OFR is mainly through three pathways: (1) The loss of energy through exhaust air of OFR; (2) The dissipation of heat
energy from the OFR chambers (OFR metal tubes) to the ambient air by convection and radiation. (3) The energy loss
through purge nitrogen between the lamps and quartz sleeves (Fig. S8).
320
Figure 6: The temperature measured by external temperature sensor at different positions inside OFR or under
different settings: (a) the measured and CFD modeled temperatures on the vertical direction at different depths; (b)
the measured and CFD modeled temperatures on the horizontal direction at different depths. All the four lamps
were turned on during measurement. The four lamps driving voltages were 5 V. The black lines and grey shadings
325
are the average ± standard deviation for temperature measured by the backplate sensor.
To clarify the energy balance of the three pathways, we measured the temperatures of the OFR tube, the purge
nitrogen, and the exhaust air. In this case, we demonstrate the results under the maximum lamp driving voltage of four UV
lamps (10 V) and at a flow rate of 5 L min-1. Figure 4 displays the measured temperature of the OFR tube as a function of
time and at different positions (inner vs external surface, bottom vs top surface). The temperature of the top external
330
surfaces increases as the gas temperature inside OFR gets higher, supporting that the OFR tube receives the heating energy
from lamps. The temperature detected at the inner surface of the OFR (40.5 °C ) is lower than the temperature of gases
(43.5 °C ), however, higher than the outside of tube surfaces (34 °C ). The lower temperature of the external surface of the
OFR than the air masses inside of the OFR is probably because i) the aluminum OFR chamber has a higher thermal mass
than the air. Although the specific heat capacity of metal (0.908 J g-1 K-1 at 301.60 K) is similar to that of air (1.005 J g-1
335
K-1 at 300 K) (Giauque and Meads, 2002; Kieffer, 1956), the flow tube is much heavier than the air due to much higher
density of the former (2700 Kg m-3 vs 1.29 Kg m-3), resulting in a lower temperature of OFR tube than the inner air; ii)
there are heating losses on the surface of OFR tube since the ambient air temperature (23 °C ) is much lower than OFR
tubes resulting in heat transfer. This heat transfer is also reflected by the decreasing temperature gradient between inner
and external OFR tube surfaces (40.5→34 °C ). We also measured the temperature of purged nitrogen at a default flow rate
340
of 0.2 L min-1. In this case, the temperature of exhausted nitrogen is ~32 °C (output), which is 9 °C higher than the
temperature of input nitrogen (23 °C ).
In addition to the air inside OFR, the temperature of exhausted air from OFR was also examined closely. A temperature
sensor was set in the Teflon Tee connector near the output of the OFR backplate. We found a slightly lower temperature
enhancement (5 °C , Fig. 5a) in the exhausted air than the values of air measured directly inside OFR, which is probably
345
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
12
influenced by the colder surfaces of the fitting or sampling lines. Using the measured temperatures shown above, we can
roughly calculate the fraction of energy loss among three pathways. Detailed information on the calculation process can be
found in Sect. S2 of the supporting information. Those model results suggest that after 105 min of heating up, 51% of the
total power (35.6 W, 8.9 W for each lamp, 4 lamps in total) during the heating process is converted into energy that leads
to the temperature increase of OFR. The remaining 49% of power is lost due to the conversion efficiency or converted into
350
other energy that does not cause the warming of airflow inside OFR. For the fraction (51%) of energy that causes warming,
the fraction of energy loss through purged nitrogen gas (at flow rate 0.2 L min-1), heated air, and OFR tubes is 0.3%, 9.8%,
and 89.9%, respectively, as shown in Fig. S9a. This suggests the heating transfer through metal tubes would be the dominant
pathway. Cooling the OFR tubes is a good way to keep the OFR near ambient temperature. As shown in the following sect.
3.5, with two fans blowing the OFR tube, the heat dissipation of the OFR tube increases, resulting in a much lower increase
355
in temperature inside the OFR.
In some OFR systems, a higher flow rate of nitrogen-purged air was applied (Zhao et al., 2021; Bruns et al., 2015; Li
et al., 2019). E.g., Li et al. (2019) set the purged nitrogen flow around 30 L min-1 for their custom-designed OFR system
with exterior lamps to keep the temperature around 25 °C . We tried to introduce a nitrogen gas flow at 20 L min-1 through
the lamps to increase the energy loss for the ARI OFR. This is almost the maximum flow rate of purged nitrogen we can
360
test due to the small inner diameter of the fitting connected to the lamp tubes. After nitrogen was injected at such a high
flow rate, we found the temperature of the lamp sleeve, the inner surface of flow tubes, and gas inside OFR dropped from
62, 40.5, and 44 °C (four lamps at 10 V) to 42, 34 and 36.5 °C (36-51% decreases in temperature increase, with ambient
temperature at 23 °C ), respectively. Based on the measured temperature, we found the energy loss fraction through purged
nitrogen increased from 0.3% to 32% (Fig. S9b). This is helpful, but it still cannot balance the heating energy input from
365
the UV lamps. The temperatures of air in the OFR (36.5°C ) are still much higher than the room temperature (23 °C ). The
large consumption of pure nitrogen gas for running such a high flow rate would also be a challenge for long-term
experiments, especially for field studies.
3.1.5 Artificially low temperature measured by primary T/RH sensor in ARI OFR
During the experiment, we found the primary T/RH sensor installed in the OFR backplate (see the sensor in Fig. S1)
370
always showed a much lower temperature (by 1-14 °C ) than the temperature sensor probing inside of OFR, as shown in
Figs. 3-6. We suspect that the lower temperatures detected by the primary sensor at the default backplate position were
mainly caused by the direct contact of this sensor with the metal backplate which is at lower temperatures. When the
primary sensor was placed around 300 mm into the OFR (closer to the lamps than the centerline), a similar temperature
was measured by the primary T/RH sensor and the one probing into the OFR (maximum difference of 1 °C ), as shown in
375
Fig. 5a. This suggests the primary T/RH sensor in the backplate with default OFR settings might lead to underestimation
of the temperature inside of OFR, which should be verified and corrected by the users based on the configuration of their
instrument.
3.2 Temperature influence on the flow field
In this section, we discuss the impact of enhanced temperature on fluid dynamics inside OFR based on Computational
380
Fluid Dynamics (CFD) simulations. Detailed setting parameters were introduced in Sect. 2.3. The temperature field where
the voltages of four UV lamps were all set to 5 V was tested here. As shown in Figs. 6a and b, the simulated temperature
distribution generally shows good agreement with the values measured directly with the temperature sensor inside OFR,
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
13
validating the reasonableness and reliability of both simulated and measured temperature distribution in OFR. In Fig. 7a,
the simulated 2-D temperature distributions display the hottest air (up to 45 °C ) wrapped around UV lamps. The rest of the
385
air parcels have a very large vertical axial temperature gradient (~10 °C ) and the air with higher temperatures between 35-
45 °C remains in the upper position of flow tubes. Such a high-temperature gradient will lead to substantial recirculation
inside OFR, based on the Richardson number calculation (Ri = 3974, which is far above 10 and indicates the existence of
turbulence) (Huang et al., 2017; Holman, 2010). Details of the Ri calculation can be found in Sect. S3 of the supporting
information.
390
To investigate the non-isothermal effect on fluid dynamics inside the OFR, the simulated flow distributions with and
without heating effects are demonstrated in Figs. 7c-j. As illustrated in Figs. 7c-d, the velocity of flow is highest after
injection through an inlet diffuser, resulting in recirculation near the edge of the walls. The other place showing a higher
velocity of air is around the output of exit (Figs. 7d and S10), which is generally used for particle output measurement. In
addition to aerosol sampling, the air for gas phase measurements was usually sampled through a perforated ring flow
395
manifold in the back of OFR to obtain more even and better-mixed air (Fig. S1). Under the typical operating condition, the
flow distribution when both gas and aerosol were sampled might be different from that when only the aerosol line was
applied. However, adding the extra output of the gas phase would greatly add to the complexity of this CFD simulation.
Thus, to simplify the simulation, we set the airflow to be only sampled through the exit port located in the center of the
OFR backplate. Moreover, when the flow ratio between gas and aerosol lines is different, the flow distribution inside OFR
400
would be changed.
When the UV light was on, the flow distribution changed substantially and had more flow bifurcation and recirculation,
as shown in Figs. 7e and 7f. There is a large recirculation in the lower part of the flow tube. A clearer demonstration of the
recirculation can be seen in Figs. 7g-j and Fig. S10. When no heat effect was introduced in the OFR, the 1-D flow profile
generally exhibited symmetry in the vertical direction (Fig. 7h). A slight distortion on the 1-D flow profile in the horizontal
405
direction is reasonable due to the influence of the random motion and pressure gradient of gas (Fig. 7g). When a vertical
axial temperature gradient exists, 1-D flow profiles are skewed due to the buoyancy of the warm air. The recirculation
effect manifested when the temperature gradient is larger, as well as when the flow rate is lower, thus can substantially
change the residence time distribution (RTD) and its average value (). can be calculated as the sum of the integrals
of signal (such as the signal of SO2 after introducing SO2 pulses) over time versus the sum of the signals (Huang et al.,
410
2017).
The average RTD values inside the OFR under heat and non-heat scenarios were also measured and simulated in Fig.
8. A 2 s pulse of 50 ppm SO2 was injected into an OFR carrier gas of 5 L min-1 and RH <10% in both the laboratory
experiment and simulation model. For laboratory measurement, N2 was used as carrier gas to prevent the reaction between
SO2 and the generated oxidant when UV lamps were turned on. Similar to the modeled results in Huang et al. (2017) and
415
the measured results in Lambe et al. (2019), the measured RTDs under higher temperatures in this study exhibit shorter
(135-145 s at 5-10 V vs 177 s at 0 V) due to the acceleration of air upon heating. The acceleration is mainly reflected
in the early arrival time and shorter tails, as shown in Fig. 8. Huang et al. (2017) found that when the flow tube is under
non-isothermal conditions even a small temperature deviation (0.2 °C ) can create secondary flow thus affecting the RTD
and substantially. The variation of the RTD when lamps were turned on and off was also simulated in the CFD model,
420
as shown in Fig. 8. Compared with the measured results, the simulated RTD generally shows an early arrival time and
broader distribution at different light settings. The simulated upon heating (173-180 s) is higher than the value when
lights were off (167 s), indicating the recirculation in the model introduced by heating prolongs the rather than
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
14
reducing the . The higher when lamps were on than off contrasts with the trend found in the measured
(Huang et al., 2017; Lambe et al., 2019). This inconsistency is probably because i) we only considered the airflow sampled
425
from the center outlet in the backplate, but not the ring flow manifold, which caused more recirculation and ii) the
recirculation was weighted more in the calculation than the real conditions, which also emphasizes the challenge of
simulating the orthogonal forces caused by the forced convection by the pressure gradient (horizontal) and buoyancy-
induced free convection (vertical) (Huang et al., 2017). Note the simulated under lamp setting of 10 V (173 s) is lower
than the under 5 V (180 s), the trend of which is consistent with the measured results, suggesting the enhanced gas
430
diffusion upon heating was properly considered except the recirculation.
Figure 7: (a) Three-dimensional simulation results demonstrating the cross-sectional temperature profiles. (b)
Simulated lateral temperature profile inside of the OFR based on CFD simulation. Panel (a) and (b) were simulated
with conditions that four lamps were set to be 5 V. (c) Three-dimensional simulation results illustrating cross-
435
sectional velocity profiles and (d) lateral velocity profiles based on CFD simulation results. Panel (c) and (d) were
simulated with conditions that lamps are off (room temperature: 25 ℃). (e-f) the same plots as panels (c) and (d)
with four lamps being set on 5 volts. Enhanced temperature influences in panels (c) and (d) were considered. One-
dimensional velocity profiles at (g) horizontal and (h) vertical directions inside the OFR at room temperature, One-
dimensional velocity profiles at (i) horizontal and (h) vertical directions under four lamps being set 5 volts.
440
In summary, the heating introduced from four lamps set to 5 V can decrease by 18-23% relative to dark OFR
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
15
experimental conditions. This is considered as an upper limit variation on for a typical OFR setting where the voltage
of lamps was usually below 3 V to obtain OH exposures less than 5 days. The variation of RTD changes the exposure time
of gas/aerosol species in the OFR, which can impact the gas and particle oxidation conditions. In the following sections,
this impact will be systematically discussed.
445
Figure 8: Residence time distribution (RTD) of SO2 inside OFR under different lamp settings. 2 s pulser of SO2 was
injected into the OFR. The average RTD values were also estimated here. The simulated results from CFD models
are shown in red lines.
3.3 Temperature influence on gas-phase reaction and OH exposure
450
In this section, the temperature influence on gas-phase reactions was systematically investigated. We take the
oxidation of SO2 inside OFR as an example to simulate its gas-phase reaction with oxidants (mainly OH) under a
temperature range from 25 °C to 40 °C (binned with 5 °C ) using the KinSim Model. The 25 °C simulation represents the
typical laboratory condition, while 40 °C was the approximate temperature observed at the upper limit of lamp voltage (5-
10 V, corresponds to photochemical age ~10-30 days, Fig. 2). In addition, to account for the RTD influences on gas-phase
455
reaction under different temperatures, the scenarios with measured RTD distributions under 25 °C and 40 °C were both
simulated here.
In general, when the temperature variation inside of OFR (the same RTD as measured at 25 °C being applied for cases
at different temperatures) was considered, the influences of temperature enhancement due to lamp heating on the gas-phase
reaction rate are very minor. The concentration of OH, HO2, O3, and O(1D) all showed a maximum of ~5% increase at the
460
highest photon flux, which is consistent with the simulated results using the same KinSim model in Li et al. (2015). The
SO2 decay and OH exposure also show negligible variations in the model (Fig. 9e and 9f). When the temperature influences
on the RTD were taken into account, the variation trend of oxidant concentration was mixed due to the combined negative
effect of the reduced average residence time () and the positive effect of temperature on oxidant concentration (Li et
al., 2015). The parameter that was most influenced is OH exposure, which shows 18-20% lower values at 40 °C than at
465
25 °C due to the shorter residence time () upon heating. Our result suggests the increase of temperature inside of OFR
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
16
due to lamp heating would have a very minor to negligible impact on the gas-phase reactions, whereas parameters related
to the RTD distribution (e.g., OH exposure) shall be considered in the current and future experiments.
Figure 9: Simulated concentrations of different oxidants from 25 to 40 °C in the OFR based on the radical
470
mechanism of the KinSim model. 800 ppb of initial SO2 and 2.2% water vapor mixing ratio (25 °C, 70% RH) were
used. The simulated results using measured RTD at 40 °C are also shown.
3.4 Temperature influence on SOA formation
The temperature can influence the SOA formation through changing gas/particle partitioning and RO2 fate, thus
affecting the SOA yield and chemical composition (Takekawa et al., 2003; Li et al., 2007; Zhang et al., 2015; Price et al.,
475
2016; Quéver et al., 2019; Kristensen et al., 2020; Atkinson et al., 1987). Here, we estimated the OA losses based on
literature results and modeling work.
For the existing OA which is sampled into the OFR, the evaporation of input OA upon heating can be roughly
estimated based on the results of thermal denuder (TD) experiments, in which the aerosols were heated from ambient/room
temperatures (20-25 °C ) to higher temperatures (typically 60-200 °C ) at a typical residence time of 10-60 s (typically ~20
480
s) (Huffman et al., 2009; Xu et al., 2020; Hu et al., 2016; Saha et al., 2017; Kolesar et al., 2015; Saha and Grieshop, 2016;
Lee et al., 2011). Previous studies suggest the evaporation of aerosol is kinetically limited, which varies with temperature,
residence time, OA mass and volatility, phase state and dissociation rates for oligomers (Cappa, 2010; Riipinen et al., 2010;
Roldin et al., 2014; Schobesberger et al., 2018). Based on the previous TD studies, the evaporation rate of 1×10-4-2.5×10-
4 per (°C ·second) was usually observed under 50 °C of TD for ambient OA (typical average mass concentration: 10-30 μg
485
m-3) (Huffman et al., 2009; Feng et al., 2023; Paciga et al., 2016). In an OFR experiment conducted on an aircraft, Nault et
al. (2018) found an average mass loss of ambient OA is ~32% due to enhanced temperature in a dark OFR compared to
the unperturbed air sampled (+17 °C on average) at a residence time of 150 s. That OFR experiment showed an evaporation
rate of 1.2×10-4 per (°C ·second), which is within the range of reported values obtained with the higher temperature TD
experiments (1×10-4-2.5×10-4 per (°C ·second)). Theoretically, when the temperature enhancement was assumed to be
490
~5 °C and a residence time of 160 s was applied, a mass loss of 8-20% was estimated for ambient OA in OFR. The enhanced
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
17
temperature inside of OFR might also impact other semi-volatile inorganic species, e.g., ammonium nitrate (Heim et al.,
2020).
For the newly formed SOA in the OFR, the temperature impact was simulated based on the SOM model (He et al.,
2022; Eluri et al., 2018). Specifically, the kinetic phase partitioning of SOA generated via OH-oxidation of typical ambient
495
VOCs was simulated (Figs. 10-11, S12-17) as a function of OFR temperature ranging from 20-40 °C (binned in 5 °C ) and
OA mass concentration from 1 to 80 μg m-3. Note that the results including SOA yield, size distribution, and O:C ratio for
n-dodecane (Figs. 10 and 11), α-pinene (Figs. S12 and S15), toluene (Figs. S13 and S16) and m-xylene (Figs. S14 and S17)
were examined under both high NOx and low NOx conditions. For dodecane, the simulated temperature-dependent SOA
under high NOx in the SOM model generally is consistent with the chamber results obtained in Lamkaddam et al. (2017).
500
Generally, the higher temperatures result in lower SOA yields due to the increased evaporation of gas-phase products
(Hildebrandt et al., 2009; Warren et al., 2009; Qi et al., 2010; Denjean et al., 2015). As shown in Figs. 10a and S12-S14,
when RTD at 25 °C was used for different temperatures, the SOA yield of different VOC species including dodecane, α-
pinene, toluene, and xylene can decrease by ~ 20% for typical temperature enhancement of 5 °C in the OFR, and by up to
40-50% at 40 °C compared to the values at typical 25 °C under high NOx condition. This decrease in SOA yield signifies
505
the substantial influences of temperature on SOA formation inside of OFR due to heating. The simulated particle size also
shows decreases as the OFR temperature goes up. The higher SOA mass and larger size of particles formed at lower
temperatures are consistent with more gas-to-particle phase partitioning, and have also been observed in various
temperature-controlled chamber studies (Clark et al., 2016; Lamkaddam et al., 2017; Boyd et al., 2017; Gao et al., 2022;
Pathak et al., 2007; Tillmann et al., 2010; Price et al., 2016; Kristensen et al., 2020). When the measured RTD at 40 °C
510
was applied in the model, an even lower SOA yield was achieved due to the shorter residence time of reactants. Under low
NOx conditions, a smaller reduction in size distribution (Fig. 11 and Figs. S15-S17) and SOA yield, which are 5-10% for
temperature enhancement of 5 °C and up to 15-35% for 15 °C as shown in Fig. 10 and Figs. S12-S14) was found compared
with high NOx conditions. The smaller reduction was mainly due to more SOA with lower volatility being formed under
low NOx conditions (e.g., acids, hydroperoxide) than the high NOx conditions (Kroll and Seinfeld, 2008; Srivastava et al.,
515
2022; Presto et al., 2005; Aruffo et al., 2022). The parameters of the SOM model were obtained based on fitting the results
to the chamber results under high and low NOx conditions, respectively (Eluri et al., 2018; Cappa et al., 2016), thus, the
parameters were set such that the volatility of SOA decreases larger as a function of the functional groups added under low
NOx condition than the high NOx condition (Cappa et al., 2013). Note that in the model we did not specifically treat the
RO2 fate e.g., a varied fraction of highly oxygenated organic molecules from autoxidation under different temperatures.
520
Using a constant yield for HOMs in the model might lead to a lower HOMs mass fraction in total SOA under high
temperatures, which would result in an underestimation of the oxidation level and an overestimation of the volatility of
SOA formed under low NOx conditions (Bianchi et al., 2019). In addition, the wall loss was corrected in the SOM model
with a constant wall loss rate of 2.5x10-3 s-1, which has been verified in the comparison of SOA model work between OFR
and chamber studies (He et al., 2022). In theory, when the temperature of the OFR tube increases, the wall loss rate of low
525
volatile organic compounds shall be smaller than under ambient temperature. Therefore, using a constant wall loss rate
here might underestimate the SOA yield.
In previous studies, to correct for heating effects on SOA yield, Chen et al. (2013) and Lambe et al. (2015b) corrected
SOA yield by 2% per K of temperature rise relative to 298 K (Stanier et al., 2008). However, this is a rough correction,
because the SOA yields of different precursors and reaction conditions are affected by temperature by different amounts.
530
For example, 0.41-0.52% per K was found for dodecane under high NOx conditions based on the SOM model while 0.87-
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
18
0.89% per K was found for α-pinene (Table S4). Note that in the OFR experiments, the decrease of SOA yield might be
even larger due to the potential mass loss of seed OA upon heating in the OFR. The simulated results from the SOM model
here provide an approximate reference to help recalibrate the SOA formation inside of OFRs. As the detailed numbers
calculated by the SOM model for different species under high and low NOx conditions can be seen in Table S4. For a
535
specific recalibration, e.g., mixed precursors, a more detailed model or temperature-controlled experiments considering the
mixing effect of precursors can be performed (Mcfiggans et al., 2019).
For the SOA chemical composition, the higher temperature inside the OFR results in increased O:C ratios of SOA,
e.g., ~0.35 (25 °C ) vs ~0.42 (40 °C ) in dodecane experiments under high NOx conditions and ~0.23 (25 °C ) vs ~0.28
(40 °C ) under low NOx condition. The enhanced O:C ratio under higher temperatures was also found in the chamber results
540
for m-xylene/NOx photooxidation in Qi et al. (2010) and O3 oxidation of α-pinene in Denjean et al. (2015). The higher O:C
under higher temperatures was probably caused by the evaporation of semi-volatile and less-oxidized components with
increased temperatures (Clark et al., 2016; Gao et al., 2022).
Figure 10: Simulated SOA yield of dodecane as a function of mass concentration of organic aerosol and temperature
545
inside OFR for (a) high NOx and (c) low NOx conditions, respectively. The simulated results using the measured
RTD obtained at 40 °C are shown as black dashed lines. The ratio of SOA yield of dodecane from different
temperatures compared to that of 25 °C under (b) high NOx and (d) low NOx conditions. The equivalent aging time
is 1 day by assuming the OH concentration is equivalent to 1.5×106 molecule cm−3 (Mao et al., 2009).
In the model, the evaporated SOA mass as a function of temperature was mainly determined by the enthalpy (
),
550
as described by the ClausiusClapeyron Equation (E.q. (1)). A varied enthalpy
dependent on saturation concentration
(C*
=-11×
) was applied as default setting based on Epstein et al. (2010). Various experimental studies using
thermodenuder for measuring OA volatility show the ambient
for OA might vary in a range of 50-150 kJ mol-1
(Epstein et al., 2010; Saha et al., 2017; Cappa and Jimenez, 2010). To further explore the sensitivity of temperature-
dependent SOA formation toward enthalpy, we demonstrate the SOA yield as a function of temperature under constant
555
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
19
values in Fig. 12 and S18. The results suggest SOA yield obtained with varied
is generally similar to that when
a constant
of 100 kJ mol-1 was applied, which is the most commonly reported
for ambient OA. Higher
suggests the SOA formation is more sensitive to the temperatures. When the upper (150 kJ mol-1) and lower limit (50 kJ
mol-1) of
are applied, the simulated results of SOA yield suggest a maximum of 22% and 90% variation, respectively,
under high NOx condition and 18% and 42% variation, respectively, for low NOx condition based on different species and
560
temperature.
Figure 11: Size distribution of dodecane under different temperatures and 30 μg m-3 organic aerosol for (a) high
NOx and (c) low NOx conditions, respectively. The O:C ratio of dodecane under different temperatures and organic
aerosol concentration under (b) high and (d) low NOx conditions. The equivalent aging time is 1 day by assuming
565
the OH concentration is equivalent to 1.5×106 molecule cm−3 (Mao et al., 2009).
Figure 12: SOA yield of dodecane as a function of temperature under different
 values (Unit: kJ mol-1). The
mass concentration of organic aerosol is assumed to be 15 μg m-3. The equivalent aging time is 1 day by assuming
the OH concentration is equivalent to 1.5×106 molecule cm−3 (Mao et al., 2009).
570
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
20
In summary, the heating effect introduced by the lamps can have an important impact on the SOA formation inside of
OFR for certain high-OH exposure applications, which shows varied extents for different precursors and reaction
conditions. The simulated results suggest that the deceased ratio of SOA output vs preexisting OA at higher OH exposures
(e.g., equivalent aging time >~5 days under low NOx conditions) in previous experiments (Hu et al., 2022; Lambe et al.,
2015a; Ortega et al., 2016; Palm et al., 2016; Saha et al., 2018) might not only due to heterogeneous reaction and/or
575
enhanced gas-phase reaction, but also probably caused by the heating evaporation. However, the yield and ambient OFR
results under equivalent aging time <~4 days shall still be valid (Palm et al., 2018). The enhanced temperature might also
impact the phase of aerosols by changing the chemical composition of OA (viscosity, O:C, etc.), as the OFR temperature
covers the usual range of glass transition temperature of ambient OA (2-87 °C ) (Li et al., 2020b; Derieux et al., 2018). The
impact of enhanced temperature on the phase state (i.e., viscosity) of aerosol phases, wall loss or other effects needs to be
580
further studied.
3.5 Approaches to reduce the heating effect
In the OFR, one approach to reducing the heating influence of the lamps is to decrease the OFR residence time, at the
expense of decreasing the maximum achievable OH exposure and time available for LVOC condensation onto aerosols
(Peng and Jimenez (2020). Another approach involves using fewer lamps at lower voltage settings; for example, using two
585
lamps at less than 3 V maintains a temperature enhancement of less than 5 °C while still achieving OH exposures of up to
5 days under low NOx conditions. In addition, it can also be useful to increase heat transfer away from the OFR, to reduce
its operating temperature e.g., blowing air with fans or air conditioners. To test this, we used two big fans (45 cm in diameter
at a distance of 30 cm from the OFR) pointed to the lower parts of the OFR. We found the maximum delta temperature
(OFR minus ambient air) for a typical high light setting decreased from 16 °C to 7 °C (44% deduction), as shown in Fig.
590
13. That test showed that blowing the flow tube with fans is a very efficient way to offset the lamps heating, and is simple
with no major trade-offs. Designing a cooling system on the outer surface of OFR by using circulating water or cold air
might also be a good way to better control the temperature inside of OFR (Watne et al., 2018; Xu and Collins, 2021; Huang
et al., 2017; Liu et al., 2018; Chu et al., 2016; Zhao et al., 2021), however, it would require a substantial redesign of the
hardware of OFR tubes and beyond the scope of this manuscript.
595
Figure 13: The measured temperature in OFR with or without fan bellowing on the flow tube in the laboratory. The
ambient temperature was also shown. Two fans were used, with a diameter of 45 cm and a rotation rate of 1400 r
min-1.
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
21
600
Figure 14: The measured temperature variation when lamp driving voltage cycles were set in a non-monotonic
pattern (10-2-6-3-5-0 V for one cycle). The red line and shading are the average temperature and standard deviation
(35.05±1.97 °C) measured at the centerline (probing distance: 300 mm) inside OFR. The purple dotted lines
represent the ambient temperature. Each lamp voltage was set for 16 min.
605
Figure 15: (a, b) The temperature variation with lamp driving voltage is explored using a monotone (10→0→10V)
and non-monotone pattern lamp power change pattern, respectively. (c) the modeled SOA formation from oxidation
of 10 ppb toluene (OM=30 μg m-3) with OH radicals as a function of OH exposure. The SOA formation under
monotone and non-monotone scenarios are both simulated. The case 3 of theoretical” indicates the temperature
were set to be constant (i.e., 25 ℃).
610
The voltage setting strategy in the OFR also can be improved. In most of the OFR laboratory and field studies, OFR
users usually changed the light setting monotonously from 0 to 10 V or from 10 to 0 V, which results in a continuously
changing temperature inside of OFR (Fig. 2, Fig. 5 and Fig. S4), therefore resulting in a variable SOA formation yield
under different temperatures (Fig. 10). To mitigate the variable heating effect, we suggest alternating between high and
low voltage settings to minimize the heat accumulation. Fig. 14 shows an example of such a light setting cycle using 10-
615
2-6-3-5-0 V. In such a way, although the average temperature inside of OFR is still higher than the ambient temperature,
the variation of measured temperature in the OFR can be kept within a narrow range and has a smaller temperature deviation
(±1.97 °C ) compared to the usual settings (Fig. 14a, ±4.76 °C ). Another advantage is that in such a way, the variation trend
of SOA yield as a function of OH exposure would have fewer uncertainties compared to those performed under
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
22
monotonically increasing or decreasing light settings. As shown in Fig. 15c, when using the measured temperature to
620
simulate the SOA formation from toluene in the OFR, the SOA showed two different trends (2-32% differences) when
monotonic and non-monotonic light settings cycles were applied. The SOA formed from the proposed non-monotonic lamp
setting cycle shows better agreement on the variation of mass concentration (Fig. 15c) and mass ratio (Fig. S19) compared
with the SOA formation at 25 °C (40.4-80.0% for non-monotone setting vs 32.9-87.3% for monotone setting), which would
introduce fewer uncertainties in the SOA yield and other aspects, e.g., chemical composition.
625
4 Conclusion
We systematically measured the temperature distribution inside of the lamp-enclosed Aerodyne PAM oxidation flow
reactor. It is found that the lamps inside of OFR create heating energy, thus leading to temperature increase inside of OFR.
A brief summary of the temperature increase and its influencing factors can be found in Table A1. The enhanced air
temperature increases as a function of the setting of lamp voltage (OH exposures) and the number of lamps applied,
630
decreases as a function of flow rate due to a shorter residence time at higher flows. The distribution of enhanced temperature
varied with special position due to the complex effects of thermal transfer and flow mixing. With default lamps from Light
Source Inc. installed, the temperature increase of air in the OFR is generally below 5 °C (at central line) when the driving
voltage of two lamps is below 3 V (typically < 5 days of equivalent atmospheric OH exposure under low-NOx conditions)
and the flow rate is 5 L min-1 (average residence time, =~150-180 s). The usage of BHK lamps typically yields smaller
635
temperature enhancement in OFR at the same OHexp due to its lower power dissipation. The heating energy loss of the OFR
system is mainly through the walls of the reaction chamber, then by exiting air inside of OFR and/or purged nitrogen.
The impact of the enhanced temperature on the flow distribution, gas, and aerosol phase chemistry inside of OFR was
systematically evaluated. The pulsed tracer measurement results suggest the enhanced temperature in OFR accelerates the
diffusion, thus resulting in a shorter average residence time (). Although box model simulation results show that the
640
temperature increase in OFR has a negligible impact on gas-phase oxidant concentrations (<~5%), however, has a certain
impact on the related parameters, e.g., OH exposure (~20% decreases results from temperature increase of 15 °C ).
The increase of temperature has larger impacts on the aerosol phase chemistry than on the gas phase. When the
enhanced temperature inside the OFR is around 5 °C at a flow rate of 5 L min-1, the evaporation loss of ambient OA was
found to be 8-20% based on thermodenuder and field experiments. The simulation from the SOM model suggests the SOA
645
yield from four typical precursors (n-dodecane, α-pinene, toluene, and m-xylene) can decrease <20% under high NOx
conditions and <10% under low NOx conditions when the temperature in OFR increases 5 °C . With the increase of
temperature, the size distributions also show a substantial decrease while increases in the O:C ratios were found. This work
demonstrates the substantial influence of temperature on SOA formation and highlights the need for consideration of the
temperature impact when using OFR as a tool for investigating aerosol chemistry. Based on the SOM model simulation
650
results in Table S4, we recommend 0.19-1.6% and 0.26-3% per K for temperature effect correction for SOA yield from
four typical precursors under high NOx and low NOx conditions, respectively.
In general, applying higher flow rate and lower lamp power, cooling the reactor with fans are recommended to keep
the temperature enhancement low in the OFR system. In addition, to control the variation of enhanced temperature
introduced by the monotonic trend for lamp settings (e.g., 0→10 V or 10→0 V), we propose to set the voltage with high
655
and low voltage alternatively (e.g., 10-2-6-3-5-0 V) to improve the heating accumulation and keep the temperature variation
within a narrow range. In summary, our evaluation of the temperature enhancement inside of lamp-enclosed OFR helps to
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
23
improve our understanding of flow distribution and chemistry inside of this OFR type, which can help to reduce the
uncertainty of OFR usage in the future.
Appendix
660
Table A1: Summary on the effect of different dimensional factors on the temperature of air inside of OFR, as well
as the impact of temperature enhancement on flow, gas and aerosol chemistry inside OFR.
Factors influence on temperature (T) of air inside OFR
Impact factors
Effect
Heating energy input
Driving voltage of lamps
Driving voltage↑ T↑
Number of lamps
NO. of lamp used T↑
Lamp types
No differences for the same brand
Lamps lasting time
Time↑ T↑ before balance
Setting voltage sequence
Set voltage monotonically T;
Set voltage monotonically usually delayed
T peak
Heating energy loss
Flow rate (corresponds to Residence
time distribution, RTD)
Flow rate (RTD ) →T
N2 purge air
Flow rate or N2 temperature T
Surrounding T of OFR
Surrounding T T
Metal tube temperature
Metal tube T T
Temperature distribution
Measured position
Closer to lamp sleeve T
Vertical
Generally, T in the upper position
Horizontal
Symmetrically distributed.
Probing depth
Probing depth→T for ARI OFR;
Probing depth→T for Penn State OFR
Temperature influences on flow, gas, and aerosol chemistry
Factors being influenced
Effect
Flow
Average RTD
T↑ RTD
Gas-phase
absolute concentration of oxidants
T↑ Minor impact (<5%)
OH exposures
T↑→ OH exposure
HOM yield
T↑→ HOM yield
Aerosol-phase
input/seed OA
T↑→ OA masses
SOA yield*
TSOA yield
Size distribution of SOA*
TPeak sizes
Oxidation level of SOA*
T↑→ Oxidation level
*More impact on high NOx regime than low NOx regime.
Data availability. The data shown in the paper are available on request from the corresponding author
(weiweihu@gig.ac.cn)
665
Author contributions. HW, AL, QH, YH, MM, SJ, YZ, ZP, YH, DD, PJ, and JJ: conceptualization, methodology,
investigation and revision. MH: simulation of flow in OFR. QH and HC: provide the solid work design of OFR. WH and
XM: supervision, project administration and funding acquisition. TP and WH wrote the paper, and all co-authors supported
the interpretation of the results and contributed to improving the paper.
Competing interests. The authors declare that they have no conflict of interest.
670
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
24
Acknowledgement
This work was supported by the National Key R&D Program of China, Young Scientist Program (No.
2021YFA1601800), the National Natural Science Foundation of China (grant No. 42375105, 42275103, 42230701),
Guangdong Pearl River Talents Program (2019QN01L948), Foundation for Program of Science and Technology Research
(2023B1212060049). ZP, MM, PJ, DD and JJ were supported by NSF Atmospheric Chemistry (AGS 2206655).
References
675
Adams, P. J. and Seinfeld, J. H.: Predicting global aerosol size distributions in general circulation models, J Geophys Res-
Atmos, 107, https://doi.org/10.1029/2001jd001010, 2002.
Ahlberg, E., Ausmeel, S., Eriksson, A., Holst, T., Karlsson, T., Brune, W. H., Frank, G., Roldin, P., Kristensson, A., and
Svenningsson, B.: No Particle Mass Enhancement from Induced Atmospheric Ageing at a Rural Site in Northern
Europe, Atmosphere, 10, 408, https://doi.org/10.3390/atmos10070408, 2019.
680
Akherati, A., Cappa, C. D., Kleeman, M. J., Docherty, K. S., Jimenez, J. L., Griffith, S. M., Dusanter, S., Stevens, P. S.,
and Jathar, S. H.: Simulating secondary organic aerosol in a regional air quality model using the statistical oxidation
model Part 3: Assessing the influence of semi-volatile and intermediate-volatility organic compounds and NOx,
Atmospheric Chemistry and Physics, 19, 4561-4594, https://doi.org/10.5194/acp-19-4561-2019, 2019.
Akherati, A., He, Y., Coggon, M. M., Koss, A. R., Hodshire, A. L., Sekimoto, K., Warneke, C., de Gouw, J., Yee, L.,
685
Seinfeld, J. H., Onasch, T. B., Herndon, S. C., Knighton, W. B., Cappa, C. D., Kleeman, M. J., Lim, C. Y., Kroll, J.
H., Pierce, J. R., and Jathar, S. H.: Oxygenated Aromatic Compounds are Important Precursors of Secondary Organic
Aerosol in Biomass-Burning Emissions, Environ Sci Technol, 54, 8568-8579, https://doi.org/10.1021/acs.est.0c01345,
2020.
Aruffo, E., Wang, J., Ye, J., Ohno, P., Qin, Y., Stewart, M., McKinney, K., Di Carlo, P., and Martin, S. T.: Partitioning of
690
Organonitrates in the Production of Secondary Organic Aerosols from alpha-Pinene Photo-Oxidation, Environ Sci
Technol, 56, 5421-5429, https://doi.org/10.1021/acs.est.1c08380, 2022.
Atkinson, R., Aschmann, S., and Winer, A.: Alkyl nitrate formation from the reaction of a series of branched RO2 radicals
with NO as a function of temperature and pressure, Journal of Atmospheric Chemistry, 5, 91-102,
https://doi.org/10.1007/bf00192505, 1987.
695
Bahreini, R., Middlebrook, A. M., Brock, C. A., de Gouw, J. A., McKeen, S. A., Williams, L. R., Daumit, K. E., Lambe,
A. T., Massoli, P., Canagaratna, M. R., Ahmadov, R., Carrasquillo, A. J., Cross, E. S., Ervens, B., Holloway, J. S.,
Hunter, J. F., Onasch, T. B., Pollack, I. B., Roberts, J. M., Ryerson, T. B., Warneke, C., Davidovits, P., Worsnop, D.
R., and Kroll, J. H.: Mass spectral analysis of organic aerosol formed downwind of the Deepwater Horizon oil spill:
field studies and laboratory confirmations, Environ Sci Technol, 46, 8025-8034, https://doi.org/10.1021/es301691k,
700
2012.
Bianchi, F., Kurten, T., Riva, M., Mohr, C., Rissanen, M. P., Roldin, P., Berndt, T., Crounse, J. D., Wennberg, P. O.,
Mentel, T. F., Wildt, J., Junninen, H., Jokinen, T., Kulmala, M., Worsnop, D. R., Thornton, J. A., Donahue, N.,
Kjaergaard, H. G., and Ehn, M.: Highly Oxygenated Organic Molecules (HOM) from Gas-Phase Autoxidation
Involving Peroxy Radicals: A Key Contributor to Atmospheric Aerosol, Chem Rev, 119, 3472-3509,
705
https://doi.org/10.1021/acs.chemrev.8b00395, 2019.
Boyd, C. M., Nah, T., Xu, L., Berkemeier, T., and Ng, N. L.: Secondary Organic Aerosol (SOA) from Nitrate Radical
Oxidation of Monoterpenes: Effects of Temperature, Dilution, and Humidity on Aerosol Formation, Mixing, and
Evaporation, Environ Sci Technol, 51, 7831-7841, https://doi.org/10.1021/acs.est.7b01460, 2017.
Bruns, E. A., El Haddad, I., Keller, A., Klein, F., Kumar, N. K., Pieber, S. M., Corbin, J. C., Slowik, J. G., Brune, W. H.,
710
Baltensperger, U., and Prévôt, A. S. H.: Inter-comparison of laboratory smog chamber and flow reactor systems on
organic aerosol yield and composition, Atmospheric Measurement Techniques, 8, 2315-2332,
https://doi.org/10.5194/amt-8-2315-2015, 2015.
Burkholder, J. B., Sander, S. P., Abbatt, J., Barker, J. R., Cappa, C., Crounse, J. D., Dibble, T. S., Huie, R. E., Kolb, C. E.,
Kurylo, M. J., Orkin, V. L., Percival, C. J., Wilmouth, D. M., and Wine, P. H.: "Chemical Kinetics and Photochemical
715
Data for Use in Atmospheric Studies, Evaluation No. 19", JPL Publication 19-5, Jet Propulsion Laboratory, Pasadena,
2019 http://jpldataeval.jpl.nasa.gov.
Cappa, C. D.: A model of aerosol evaporation kinetics in a thermodenuder, Atmospheric Measurement Techniques, 3, 579-
592, https://doi.org/10.5194/amt-3-579-2010, 2010.
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
25
Cappa, C. D. and Jimenez, J. L.: Quantitative estimates of the volatility of ambient organic aerosol, Atmospheric Chemistry
720
and Physics, 10, 5409-5424, https://doi.org/10.5194/acp-10-5409-2010, 2010.
Cappa, C. D. and Wilson, K. R.: Multi-generation gas-phase oxidation, equilibrium partitioning, and the formation and
evolution of secondary organic aerosol, Atmospheric Chemistry and Physics, 12, 9505-9528,
https://doi.org/10.5194/acp-12-9505-2012, 2012.
Cappa, C. D., Zhang, X., Loza, C. L., Craven, J. S., Yee, L. D., and Seinfeld, J. H.: Application of the Statistical Oxidation
725
Model (SOM) to Secondary Organic Aerosol formation from photooxidation of of C12 alkanes, Atmospheric
Chemistry and Physics, 13, 1591-1606, https://doi.org/10.5194/acp-13-1591-2013, 2013.
Cappa, C. D., Jathar, S. H., Kleeman, M. J., Docherty, K. S., Jimenez, J. L., Seinfeld, J. H., and Wexler, A. S.: Simulating
secondary organic aerosol in a regional air quality model using the statistical oxidation model Part 2: Assessing the
influence of vapor wall losses, Atmospheric Chemistry and Physics, 16, 3041-3059, https://doi.org/10.5194/acp-16-
730
3041-2016, 2016.
Carter, W., Cockeriii, D., Fitz, D., Malkina, I., Bumiller, K., Sauer, C., Pisano, J., Bufalino, C., and Song, C.: A new
environmental chamber for evaluation of gas-phase chemical mechanisms and secondary aerosol formation,
Atmospheric Environment, 39, 7768-7788, https://doi.org/10.1016/j.atmosenv.2005.08.040, 2005.
Charnawskas, J. C., Alpert, P. A., Lambe, A. T., Berkemeier, T., O'Brien, R. E., Massoli, P., Onasch, T. B., Shiraiwa, M.,
735
Moffet, R. C., Gilles, M. K., Davidovits, P., Worsnop, D. R., and Knopf, D. A.: Condensed-phase biogenic-
anthropogenic interactions with implications for cold cloud formation, Faraday Discuss, 200, 165-194,
https://doi.org/10.1039/c7fd00010c, 2017.
Chen, S., Brune, W. H., Lambe, A. T., Davidovits, P., and Onasch, T. B.: Modeling organic aerosol from the oxidation of
α-pinene in a Potential Aerosol Mass (PAM) chamber, Atmospheric Chemistry and Physics, 13, 5017-5031,
740
https://doi.org/10.5194/acp-13-5017-2013, 2013.
Chen, W., Ye, Y., Hu, W., Zhou, H., Pan, T., Wang, Y., Song, W., Song, Q., Ye, C., Wang, C., Wang, B., Huang, S., Yuan,
B., Zhu, M., Lian, X., Zhang, G., Bi, X., Jiang, F., Liu, J., Canonaco, F., Prevot, A. S. H., Shao, M., and Wang, X.:
RealTime Characterization of Aerosol Compositions, Sources, and Aging Processes in Guangzhou During PRIDE
GBA 2018 Campaign, Journal of Geophysical Research: Atmospheres, 126, e2021JD035114,
745
https://doi.org/10.1029/2021jd035114, 2021.
Cheng, X., Chen, Q., Jie Li, Y., Zheng, Y., Liao, K., and Huang, G.: Highly oxygenated organic molecules produced by
the oxidation of benzene and toluene in a wide range of OH exposure and
NO&lt;sub&gt;&lt;i&gt;x&lt;/i&gt;&lt;/sub&gt; conditions, Atmospheric Chemistry and Physics, 21, 12005-12019,
https://doi.org/10.5194/acp-21-12005-2021, 2021.
750
Chu, B., Liu, Y., Ma, Q., Ma, J., He, H., Wang, G., Cheng, S., and Wang, X.: Distinct potential aerosol masses under
different scenarios of transport at a suburban site of Beijing, J Environ Sci (China), 39, 52-61,
https://doi.org/10.1016/j.jes.2015.11.003, 2016.
Clark, C. H., Kacarab, M., Nakao, S., Asa-Awuku, A., Sato, K., and Cocker, D. R., 3rd: Temperature Effects on Secondary
Organic Aerosol (SOA) from the Dark Ozonolysis and Photo-Oxidation of Isoprene, Environ Sci Technol, 50, 5564-
755
5571, https://doi.org/10.1021/acs.est.5b05524, 2016.
Cocker, D. R., 3rd, Flagan, R. C., and Seinfeld, J. H.: State-of-the-art chamber facility for studying atmospheric aerosol
chemistry, Environ Sci Technol, 35, 2594-2601, https://doi.org/10.1021/es0019169, 2001.
Cooper, P. L. and Abbatt, J. P. D.: Heterogeneous Interactions of OH and HO2 Radicals with Surfaces Characteristic of
Atmospheric Particulate Matter, The Journal of Physical Chemistry, 100, 2249-2254,
760
https://doi.org/10.1021/jp952142z, 1996.
Denjean, C., Formenti, P., Picquet-Varrault, B., Camredon, M., Pangui, E., Zapf, P., Katrib, Y., Giorio, C., Tapparo, A.,
Temime-Roussel, B., Monod, A., Aumont, B., and Doussin, J. F.: Aging of secondary organic aerosol generated from
the ozonolysis of α-pinene: effects of ozone, light and temperature, Atmospheric Chemistry and Physics, 15, 883-897,
https://doi.org/10.5194/acp-15-883-2015, 2015.
765
DeRieux, W.-S. W., Li, Y., Lin, P., Laskin, J., Laskin, A., Bertram, A. K., Nizkorodov, S. A., and Shiraiwa, M.: Predicting
the glass transition temperature and viscosity of secondary organic material using molecular composition,
Atmospheric Chemistry and Physics, 18, 6331-6351, https://doi.org/10.5194/acp-18-6331-2018, 2018.
Donahue, N. M., Robinson, A. L., Stanier, C. O., and Pandis, S. N.: Coupled partitioning, dilution, and chemical aging of
semivolatile organics, Environ Sci Technol, 40, 2635-2643, https://doi.org/10.1021/es052297c, 2006.
770
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
26
Eluri, S., Cappa, C. D., Friedman, B., Farmer, D. K., and Jathar, S. H.: Modeling the formation and composition of
secondary organic aerosol from diesel exhaust using parameterized and semi-explicit chemistry and thermodynamic
models, Atmospheric Chemistry and Physics, 18, 13813-13838, https://doi.org/10.5194/acp-18-13813-2018, 2018.
Epstein, S. A., Riipinen, I., and Donahue, N. M.: A semiempirical correlation between enthalpy of vaporization and
saturation concentration for organic aerosol, Environ Sci Technol, 44, 743-748, https://doi.org/10.1021/es902497z,
775
2010.
Ezell, M. J., Johnson, S. N., Yu, Y., Perraud, V., Bruns, E. A., Alexander, M. L., Zelenyuk, A., Dabdub, D., and Finlayson-
Pitts, B. J.: A New Aerosol Flow System for Photochemical and Thermal Studies of Tropospheric Aerosols, Aerosol
Science and Technology, 44, 329-338, https://doi.org/10.1080/02786821003639700, 2010.
Feng, S., Gao, D., Liao, F., Zhou, F., and Wang, X.: The health effects of ambient PM2.5 and potential mechanisms,
780
Ecotoxicol Environ Saf, 128, 67-74, https://doi.org/10.1016/j.ecoenv.2016.01.030, 2016.
Feng, T., Wang, Y., Hu, W., Zhu, M., Song, W., Chen, W., Sang, Y., Fang, Z., Deng, W., Fang, H., Yu, X., Wu, C., Yuan,
B., Huang, S., Shao, M., Huang, X., He, L., Lee, Y. R., Huey, L. G., Canonaco, F., Prevot, A. S. H., and Wang, X.:
Impact of aging on the sources, volatility, and viscosity of organic aerosols in Chinese outflows, Atmos. Chem. Phys.,
23, 611-636, https://doi.org/10.5194/acp-23-611-2023, 2023.
785
Gao, L., Song, J., Mohr, C., Huang, W., Vallon, M., Jiang, F., Leisner, T., and Saathoff, H.: Kinetics, SOA yields and
chemical composition of secondary organic aerosol from β-caryophyllene ozonolysis with and without nitrogen oxides
between 213 and 313 K, Atmospheric Chemistry and Physics, 22, 6001-6020, https://doi.org/10.5194/acp-22-6001-
2022, 2022.
George, I. J., Vlasenko, A., Slowik, J. G., Broekhuizen, K., and Abbatt, J. P. D.: Heterogeneous oxidation of saturated
790
organic aerosols by hydroxyl radicals: uptake kinetics, condensed-phase products, and particle size change,
Atmospheric Chemistry and Physics, 7, 4187-4201, https://doi.org/10.5194/acp-7-4187-2007, 2007.
Ghia, U., Ghia, K. N., and Shin, C. T.: High-Resolutions for incompressible flow using the Navier-Stokes equations and a
multigrid method, Journal of Computational Physics, 48, 387-411, https://doi.org/10.1016/0021-9991(82)90058-4,
1982.
795
Giauque, W. F. and Meads, P. F.: The Heat Capacities and Entropies of Aluminum and Copper from 15 to 300°K, Journal
of the American Chemical Society, 63, 1897-1901, https://doi.org/10.1021/ja01852a027, 2002.
Hallquist, M., Wenger, J. C., Baltensperger, U., Rudich, Y., Simpson, D., Claeys, M., Dommen, J., Donahue, N. M., George,
C., Goldstein, A. H., Hamilton, J. F., Herrmann, H., Hoffmann, T., Iinuma, Y., Jang, M., Jenkin, M. E., Jimenez, J.
L., Kiendler-Scharr, A., Maenhaut, W., McFiggans, G., Mentel, T. F., Monod, A., Prévôt, A. S. H., Seinfeld, J. H.,
800
Surratt, J. D., Szmigielski, R., and Wildt, J.: The formation, properties and impact of secondary organic aerosol:
current and emerging issues, Atmospheric Chemistry and Physics, 9, 5155-5236, https://doi.org/10.5194/acp-9-5155-
2009, 2009.
Hanson, D. R. and Lovejoy, E. R.: The Reaction of CIONO2 with Submicrometer Sulfuric Acid Aerosol, Science, 267,
1326-1328, https://doi.org/10.1126/science.267.5202.1326, 1995.
805
He, Y., Lambe, A. T., Seinfeld, J. H., Cappa, C. D., Pierce, J. R., and Jathar, S. H.: Process-Level Modeling Can
Simultaneously Explain Secondary Organic Aerosol Evolution in Chambers and Flow Reactors, Environ Sci Technol,
56, 6262-6273, https://doi.org/10.1021/acs.est.1c08520, 2022.
He, Y., Akherati, A., Nah, T., Ng, N. L., Garofalo, L. A., Farmer, D. K., Shiraiwa, M., Zaveri, R. A., Cappa, C. D., Pierce,
J. R., and Jathar, S. H.: Particle Size Distribution Dynamics Can Help Constrain the Phase State of Secondary Organic
810
Aerosol, Environ Sci Technol, 55, 1466-1476, https://doi.org/10.1021/acs.est.0c05796, 2021.
Heim, E. W., Dibb, J., Scheuer, E., Jost, P. C., Nault, B. A., Jimenez, J. L., Peterson, D., Knote, C., Fenn, M., Hair, J.,
Beyersdorf, A. J., Corr, C., and Anderson, B. E.: Asian dust observed during KORUS-AQ facilitates the uptake and
incorporation of soluble pollutants during transport to South Korea, Atmospheric Environment, 224,
https://doi.org/10.1016/j.atmosenv.2020.117305, 2020.
815
Hildebrandt, L., Donahue, N. M., and Pandis, S. N.: High formation of secondary organic aerosol from the photo-oxidation
of toluene, Atmospheric Chemistry and Physics, 9, 2973-2986, https://doi.org/10.5194/acp-9-2973-2009, 2009.
Hodshire, A. L., Palm, B. B., Alexander, M. L., Bian, Q., Campuzano-Jost, P., Cross, E. S., Day, D. A., de , S. S.,
Guenther, A. B., Hansel, A., Hunter, J. F., Jud, W., Karl, T., Kim, S., Kroll, J. H., Park, J.-H., Peng, Z., Seco, R.,
Smith, J. N., Jimenez, J. L., and Pierce, J. R.: Constraining nucleation, condensation, and chemistry in oxidation flow
820
reactors using size-distribution measurements and aerosol microphysical modeling, Atmospheric Chemistry and
Physics, 18, 12433-12460, https://doi.org/10.5194/acp-18-12433-2018, 2018.
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
27
Holman, J. P.: Heat Transfer, 10 edn, McGraw-Hill, New York, USA, 2010.
Hu, W., Palm, B. B., Day, D. A., Campuzano-Jost, P., Krechmer, J. E., Peng, Z., de , S. S., Martin, S. T., Alexander, M.
L., Baumann, K., Hacker, L., Kiendler-Scharr, A., Koss, A. R., de Gouw, J. A., Goldstein, A. H., Seco, R., Sjostedt,
825
S. J., Park, J.-H., Guenther, A. B., Kim, S., Canonaco, F., Prévôt, A. S. H., Brune, W. H., and Jimenez, J. L.: Volatility
and lifetime against OH heterogeneous reaction of ambient isoprene-epoxydiols-derived secondary organic aerosol
(IEPOX-SOA), Atmospheric Chemistry and Physics, 16, 11563-11580, https://doi.org/10.5194/acp-16-11563-2016,
2016.
Hu, W., Zhou, H., Chen, W., Ye, Y., Pan, T., Wang, Y., Song, W., Zhang, H., Deng, W., Zhu, M., Wang, C., Wu, C., Ye,
830
C., Wang, Z., Yuan, B., Huang, S., Shao, M., Peng, Z., Day, D. A., Campuzano-Jost, P., Lambe, A. T., Worsnop, D.
R., Jimenez, J. L., and Wang, X.: Oxidation Flow Reactor Results in a Chinese Megacity Emphasize the Important
Contribution of S/IVOCs to Ambient SOA Formation, Environ Sci Technol, 56, 6880-6893,
https://doi.org/10.1021/acs.est.1c03155, 2022.
Huang, R. J., Zhang, Y., Bozzetti, C., Ho, K. F., Cao, J. J., Han, Y., Daellenbach, K. R., Slowik, J. G., Platt, S. M., Canonaco,
835
F., Zotter, P., Wolf, R., Pieber, S. M., Bruns, E. A., Crippa, M., Ciarelli, G., Piazzalunga, A., Schwikowski, M.,
Abbaszade, G., Schnelle-Kreis, J., Zimmermann, R., An, Z., Szidat, S., Baltensperger, U., El Haddad, I., and Prevot,
A. S.: High secondary aerosol contribution to particulate pollution during haze events in China, Nature, 514, 218-222,
https://doi.org/10.1038/nature13774, 2014.
Huang, Y., Coggon, M. M., Zhao, R., Lignell, H., Bauer, M. U., Flagan, R. C., and Seinfeld, J. H.: The Caltech
840
Photooxidation Flow Tube reactor: design, fluid dynamics and characterization, Atmospheric Measurement
Techniques, 10, 839-867, https://doi.org/10.5194/amt-10-839-2017, 2017.
Huffman, J. A., Docherty, K. S., Aiken, A. C., Cubison, M. J., Ulbrich, I. M., DeCarlo, P. F., Sueper, D., Jayne, J. T.,
Worsnop, D. R., Ziemann, P. J., and Jimenez, J. L.: Chemically-resolved aerosol volatility measurements from two
megacity field studies, Atmospheric Chemistry and Physics, 9, 7161-7182, https://doi.org/10.5194/acp-9-7161-2009,
845
2009.
Ihalainen, M., Tiitta, P., Czech, H., Yli-Pirilä, P., Hartikainen, A., Kortelainen, M., Tissari, J., Stengel, B., Sklorz, M.,
Suhonen, H., Lamberg, H., Leskinen, A., Kiendler-Scharr, A., Harndorf, H., Zimmermann, R., Jokiniemi, J., and
Sippula, O.: A novel high-volume Photochemical Emission Aging flow tube Reactor (PEAR), Aerosol Science and
Technology, 53, 276-294, https://doi.org/10.1080/02786826.2018.1559918, 2019.
850
Jathar, S. H., Cappa, C. D., Wexler, A. S., Seinfeld, J. H., and Kleeman, M. J.: Multi-generational oxidation model to
simulate secondary organic aerosol in a 3-D air quality model, Geoscientific Model Development, 8, 2553-2567,
https://doi.org/10.5194/gmd-8-2553-2015, 2015.
Jathar, S. H., Cappa, C. D., Wexler, A. S., Seinfeld, J. H., and Kleeman, M. J.: Simulating secondary organic aerosol in a
regional air quality model using the statistical oxidation model Part 1: Assessing the influence of constrained multi-
855
generational ageing, Atmospheric Chemistry and Physics, 16, 2309-2322, https://doi.org/10.5194/acp-16-2309-2016,
2016.
Jensen, L. N., Canagaratna, M. R., Kristensen, K., Quéver, L. L. J., Rosati, B., Teiwes, R., Glasius, M., Pedersen, H. B.,
Ehn, M., and Bilde, M.: Temperature and volatile organic compound concentrations as controlling factors for chemical
composition of α-pinene-derived secondary organic aerosol, Atmospheric Chemistry and Physics, 21, 11545-11562,
860
https://doi.org/10.5194/acp-21-11545-2021, 2021.
Jimenez, J. L., Canagaratna, M. R., Donahue, N. M., Prevot, A. S., Zhang, Q., Kroll, J. H., DeCarlo, P. F., Allan, J. D.,
Coe, H., Ng, N. L., Aiken, A. C., Docherty, K. S., Ulbrich, I. M., Grieshop, A. P., Robinson, A. L., Duplissy, J., Smith,
J. D., Wilson, K. R., Lanz, V. A., Hueglin, C., Sun, Y. L., Tian, J., Laaksonen, A., Raatikainen, T., Rautiainen, J.,
Vaattovaara, P., Ehn, M., Kulmala, M., Tomlinson, J. M., Collins, D. R., Cubison, M. J., Dunlea, E. J., Huffman, J.
865
A., Onasch, T. B., Alfarra, M. R., Williams, P. I., Bower, K., Kondo, Y., Schneider, J., Drewnick, F., Borrmann, S.,
Weimer, S., Demerjian, K., Salcedo, D., Cottrell, L., Griffin, R., Takami, A., Miyoshi, T., Hatakeyama, S., Shimono,
A., Sun, J. Y., Zhang, Y. M., Dzepina, K., Kimmel, J. R., Sueper, D., Jayne, J. T., Herndon, S. C., Trimborn, A. M.,
Williams, L. R., Wood, E. C., Middlebrook, A. M., Kolb, C. E., Baltensperger, U., and Worsnop, D. R.: Evolution of
organic aerosols in the atmosphere, Science, 326, 1525-1529, https://doi.org/10.1126/science.1180353, 2009.
870
Kang, E., Root, M. J., Toohey, D. W., and Brune, W. H.: Introducing the concept of Potential Aerosol Mass (PAM),
Atmospheric Chemistry and Physics, 7, 5727-5744, https://doi.org/10.5194/acp-7-5727-2007, 2007.
Kang, E., Lee, M., Brune, W. H., Lee, T., Park, T., Ahn, J., and Shang, X.: Photochemical aging of aerosol particles in
different air masses arriving at Baengnyeong Island, Korea, Atmospheric Chemistry and Physics, 18, 6661-6677,
https://doi.org/10.5194/acp-18-6661-2018, 2018.
875
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
28
Katrib, Y., Biskos, G., Buseck, P. R., Davidovits, P., Jayne, J. T., Mochida, M., Wise, M. E., Worsnop, D. R., and Martin,
S. T.: Ozonolysis of mixed oleic-acid/stearic-acid particles: reaction kinetics and chemical morphology, J Phys Chem
A, 109, 10910-10919, https://doi.org/10.1021/jp054714d, 2005.
Keller, A., Kalbermatter, D. M., Wolfer, K., Specht, P., Steigmeier, P., Resch, J., Kalberer, M., Hammer, T., and Vasilatou,
K.: The organic coating unit, an all-in-one system for reproducible generation of secondary organic matter aerosol,
880
Aerosol Science and Technology, 56, 947-958, https://doi.org/10.1080/02786826.2022.2110448, 2022.
Khalaj, F., Rivas-Ubach, A., Anderton, C. R., China, S., Mooney, K., and Faiola, C. L.: Acyclic Terpenes Reduce
Secondary Organic Aerosol Formation from Emissions of a Riparian Shrub, ACS Earth and Space Chemistry, 5, 1242-
1253, https://doi.org/10.1021/acsearthspacechem.0c00300, 2021.
Kieffer, W. F.: Tables of thermal properties of gases, Journal of Chemical Education, 33, A477,
885
https://doi.org/10.1021/ed033pA477.1, 1956.
Klyta, J. and Czaplicka, M.: Determination of secondary organic aerosol in particulate matter Short review,
Microchemical Journal, 157, https://doi.org/10.1016/j.microc.2020.104997, 2020.
Knopf, D. A., Anthony, L. M., and Bertram, A. K.: Reactive uptake of O3 by multicomponent and multiphase mixtures
containing oleic acid, J Phys Chem A, 109, 5579-5589, https://doi.org/10.1021/jp0512513, 2005.
890
Kolesar, K. R., Chen, C., Johnson, D., and Cappa, C. D.: The influences of mass loading and rapid dilution of secondary
organic aerosol on particle volatility, Atmospheric Chemistry and Physics, 15, 9327-9343,
https://doi.org/10.5194/acp-15-9327-2015, 2015.
Kramer, A. L., Suski, K. J., Bell, D. M., Zelenyuk, A., and Massey Simonich, S. L.: Formation of Polycyclic Aromatic
Hydrocarbon Oxidation Products in alpha-Pinene Secondary Organic Aerosol Particles Formed through Ozonolysis,
895
Environ Sci Technol, 53, 6669-6677, https://doi.org/10.1021/acs.est.9b01732, 2019.
Kristensen, K., Jensen, L. N., Quéver, L. L. J., Christiansen, S., Rosati, B., Elm, J., Teiwes, R., Pedersen, H. B., Glasius,
M., Ehn, M., and Bilde, M.: The Aarhus Chamber Campaign on Highly Oxygenated Organic Molecules and Aerosols
(ACCHA): particle formation, organic acids, and dimer esters from α-pinene ozonolysis at different temperatures,
Atmospheric Chemistry and Physics, 20, 12549-12567, https://doi.org/10.5194/acp-20-12549-2020, 2020.
900
Kroll, J. H. and Seinfeld, J. H.: Chemistry of secondary organic aerosol: Formation and evolution of low-volatility organics
in the atmosphere, Atmospheric Environment, 42, 3593-3624, https://doi.org/10.1016/j.atmosenv.2008.01.003, 2008.
Lambe, A., Massoli, P., Zhang, X., Canagaratna, M., Nowak, J., Daube, C., Yan, C., Nie, W., Onasch, T., Jayne, J., Kolb,
C., Davidovits, P., Worsnop, D., and Brune, W.: Controlled nitric oxide production via O(1D)+N2O reactions for use
in oxidation flow reactor studies, Atmospheric Measurement Techniques, 10, 2283-2298, https://doi.org/10.5194/amt-
905
10-2283-2017, 2017.
Lambe, A. T., Ahern, A. T., Wright, J. P., Croasdale, D. R., Davidovits, P., and Onasch, T. B.: Oxidative aging and cloud
condensation nuclei activation of laboratory combustion soot, Journal of Aerosol Science, 79, 31-39,
https://doi.org/10.1016/j.jaerosci.2014.10.001, 2015a.
Lambe, A. T., Krechmer, J. E., Peng, Z., Casar, J. R., Carrasquillo, A. J., Raff, J. D., Jimenez, J. L., and Worsnop, D. R.:
910
HOx and NOx production in oxidation flow reactors via photolysis of isopropyl nitrite, isopropyl nitrite-d7, and 1,3-
propyl dinitrite at λ = 254, 350, and 369nm, Atmospheric Measurement Techniques, 12, 299-311,
https://doi.org/10.5194/amt-12-299-2019, 2019.
Lambe, A. T., Onasch, T. B., Massoli, P., Croasdale, D. R., Wright, J. P., Ahern, A. T., Williams, L. R., Worsnop, D. R.,
Brune, W. H., and Davidovits, P.: Laboratory studies of the chemical composition and cloud condensation nuclei
915
(CCN) activity of secondary organic aerosol (SOA) and oxidized primary organic aerosol (OPOA), Atmospheric
Chemistry and Physics, 11, 8913-8928, https://doi.org/10.5194/acp-11-8913-2011, 2011a.
Lambe, A. T., Chhabra, P. S., Onasch, T. B., Brune, W. H., Hunter, J. F., Kroll, J. H., Cummings, M. J., Brogan, J. F.,
Parmar, Y., Worsnop, D. R., Kolb, C. E., and Davidovits, P.: Effect of oxidant concentration, exposure time, and seed
particles on secondary organic aerosol chemical composition and yield, Atmospheric Chemistry and Physics, 15,
920
3063-3075, https://doi.org/10.5194/acp-15-3063-2015, 2015b.
Lambe, A. T., Ahern, A. T., Williams, L. R., Slowik, J. G., Wong, J. P. S., Abbatt, J. P. D., Brune, W. H., Ng, N. L., Wright,
J. P., Croasdale, D. R., Worsnop, D. R., Davidovits, P., and Onasch, T. B.: Characterization of aerosol photooxidation
flow reactors: heterogeneous oxidation, secondary organic aerosol formation and cloud condensation nuclei activity
measurements, Atmospheric Measurement Techniques, 4, 445-461, https://doi.org/10.5194/amt-4-445-2011, 2011b.
925
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
29
Lamkaddam, H., Gratien, A., Pangui, E., Cazaunau, M., Picquet-Varrault, B., and Doussin, J. F.: High-NOx Photooxidation
of n-Dodecane: Temperature Dependence of SOA Formation, Environ Sci Technol, 51, 192-201,
https://doi.org/10.1021/acs.est.6b03821, 2017.
Lee, B.-H., Pierce, J. R., Engelhart, G. J., and Pandis, S. N.: Volatility of secondary organic aerosol from the ozonolysis of
monoterpenes, Atmospheric Environment, 45, 2443-2452, https://doi.org/10.1016/j.atmosenv.2011.02.004, 2011.
930
Lei, Z., Olson, N. E., Zhang, Y., Chen, Y., Lambe, A. T., Zhang, J., White, N. J., Atkin, J. M., Banaszak Holl, M. M.,
Zhang, Z., Gold, A., Surratt, J. D., and Ault, A. P.: Morphology and Viscosity Changes after Reactive Uptake of
Isoprene Epoxydiols in Submicrometer Phase Separated Particles with Secondary Organic Aerosol Formed from
Different Volatile Organic Compounds, ACS Earth and Space Chemistry, 6, 871-882,
https://doi.org/10.1021/acsearthspacechem.1c00156, 2022.
935
Li, J., Wang, W., Li, K., Zhang, W., Peng, C., Zhou, L., Shi, B., Chen, Y., Liu, M., Li, H., and Ge, M.: Temperature effects
on optical properties and chemical composition of secondary organic aerosol derived from n-dodecane, Atmospheric
Chemistry and Physics, 20, 8123-8137, https://doi.org/10.5194/acp-20-8123-2020, 2020a.
Li, K., Liggio, J., Lee, P., Han, C., Liu, Q., and Li, S.-M.: Secondary organic aerosol formation from α-pinene, alkanes,
and oil-sands-related precursors in a new oxidation flow reactor, Atmospheric Chemistry and Physics, 19, 9715-9731,
940
https://doi.org/10.5194/acp-19-9715-2019, 2019.
Li, Q., Hu, D., Leungsakul, S., and Kamens, R. M.: Large outdoor chamber experiments and computer simulations: (I)
Secondary organic aerosol formation from the oxidation of a mixture of d-limonene and α-pinene, Atmospheric
Environment, 41, 9341-9352, https://doi.org/10.1016/j.atmosenv.2007.09.017, 2007.
Li, R., Palm, B. B., Ortega, A. M., Hlywiak, J., Hu, W., Peng, Z., Day, D. A., Knote, C., Brune, W. H., de Gouw, J. A.,
945
and Jimenez, J. L.: Modeling the radical chemistry in an oxidation flow reactor: radical formation and recycling,
sensitivities, and the OH exposure estimation equation, J Phys Chem A, 119, 4418-4432,
https://doi.org/10.1021/jp509534k, 2015.
Li, Y., Day, D. A., Stark, H., Jimenez, J. L., and Shiraiwa, M.: Predictions of the glass transition temperature and viscosity
of organic aerosols from volatility distributions, Atmospheric Chemistry and Physics, 20, 8103-8122,
950
https://doi.org/10.5194/acp-20-8103-2020, 2020b.
Liao, K., Chen, Q., Liu, Y., Li, Y. J., Lambe, A. T., Zhu, T., Huang, R. J., Zheng, Y., Cheng, X., Miao, R., Huang, G.,
Khuzestani, R. B., and Jia, T.: Secondary Organic Aerosol Formation of Fleet Vehicle Emissions in China: Potential
Seasonality of Spatial Distributions, Environ Sci Technol, 55, 7276-7286, https://doi.org/10.1021/acs.est.0c08591,
2021.
955
Link, M. F., Friedman, B., Fulgham, R., Brophy, P., Galang, A., Jathar, S. H., Veres, P., Roberts, J. M., and Farmer, D. K.:
Photochemical processing of diesel fuel emissions as a large secondary source of isocyanic acid (HNCO), Geophysical
Research Letters, 43, 4033-4041, https://doi.org/10.1002/2016gl068207, 2016.
Liu, J., Chu, B., Chen, T., Liu, C., Wang, L., Bao, X., and He, H.: Secondary Organic Aerosol Formation from Ambient
Air at an Urban Site in Beijing: Effects of OH Exposure and Precursor Concentrations, Environ Sci Technol, 52, 6834-
960
6841, https://doi.org/10.1021/acs.est.7b05701, 2018.
Liu, T., Zhou, L., Liu, Q., Lee, B. P., Yao, D., Lu, H., Lyu, X., Guo, H., and Chan, C. K.: Secondary Organic Aerosol
Formation from Urban Roadside Air in Hong Kong, Environ Sci Technol, 53, 3001-3009,
https://doi.org/10.1021/acs.est.8b06587, 2019.
Louvaris, E. E., Karnezi, E., Kostenidou, E., Kaltsonoudis, C., and Pandis, S. N.: Estimation of the volatility distribution
965
of organic aerosol combining thermodenuder and isothermal dilution measurements, Atmospheric Measurement
Techniques, 10, 3909-3918, https://doi.org/10.5194/amt-10-3909-2017, 2017.
Mao, J., Ren, X., Brune, W. H., Olson, J. R., Crawford, J. H., Fried, A., Huey, L. G., Cohen, R. C., Heikes, B., Singh, H.
B., Blake, D. R., Sachse, G. W., Diskin, G. S., Hall, S. R., and Shetter, R. E.: Airborne measurement of OH reactivity
during INTEX-B, Atmospheric Chemistry and Physics, 9, 163-173, https://doi.org/10.5194/acp-9-163-2009, 2009.
970
Martin-Reviejo, M. and Wirtz, K.: Is benzene a precursor for secondary organic aerosol?, Environ Sci Technol, 39, 1045-
1054, https://doi.org/10.1021/es049802a, 2005.
McFiggans, G., Mentel, T. F., Wildt, J., Pullinen, I., Kang, S., Kleist, E., Schmitt, S., Springer, M., Tillmann, R., Wu, C.,
Zhao, D., Hallquist, M., Faxon, C., Le Breton, M., Hallquist, A. M., Simpson, D., Bergstrom, R., Jenkin, M. E., Ehn,
M., Thornton, J. A., Alfarra, M. R., Bannan, T. J., Percival, C. J., Priestley, M., Topping, D., and Kiendler-Scharr, A.:
975
Secondary organic aerosol reduced by mixture of atmospheric vapours, Nature, 565, 587-593,
https://doi.org/10.1038/s41586-018-0871-y, 2019.
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
30
Mitroo, D., Sun, Y., Combest, D. P., Kumar, P., and Williams, B. J.: Assessing the degree of plug flow in oxidation flow
reactors (OFRs): a study on a potential aerosol mass (PAM) reactor, Atmospheric Measurement Techniques, 11, 1741-
1756, https://doi.org/10.5194/amt-11-1741-2018, 2018.
980
Murschell, T. and Farmer, D. K.: Atmospheric OH Oxidation of Three Chlorinated Aromatic Herbicides, Environ Sci
Technol, 52, 4583-4591, https://doi.org/10.1021/acs.est.7b06025, 2018.
Myhre, G., Shindell, D., Bréon, F.-M., Collins, W., Fuglestvedt, J., Huang, J., Koch, D., Lamarque, J.-F., Lee, D., Mendoza,
B., Nakajima, T., Robock, A., Stephens, G., Takemura, T., and Zhang, H.: Anthropogenic and Natural Radiative
Forcing, in: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth
985
Assessment Report of the Intergovernmental Panel on Climate Change, edited by: Stocker, T. F., Qin, D., Plattner,
G.-K., Tignor, M., Allen, S. K., Boschung, J., Nauels, A., Xia, Y., Bex, V., and Midgley, P. M., p. 659, Cambridge
University Press, Cambridge, United Kingdom and New York, NY, USA, United Kingdom and New York, NY, USA,
available at: https://www.ipcc.ch/report/ar5/wg1/ (last access: 6 December 2017), 2013.
Nault, B. A., Campuzano-Jost, P., Day, D. A., Schroder, J. C., Anderson, B., Beyersdorf, A. J., Blake, D. R., Brune, W. H.,
990
Choi, Y., Corr, C. A., de Gouw, J. A., Dibb, J., DiGangi, J. P., Diskin, G. S., Fried, A., Huey, L. G., Kim, M. J., Knote,
C. J., Lamb, K. D., Lee, T., Park, T., Pusede, S. E., Scheuer, E., Thornhill, K. L., Woo, J.-H., and Jimenez, J. L.:
Secondary organic aerosol production from local emissions dominates the organic aerosol budget over Seoul, South
Korea, during KORUS-AQ, Atmospheric Chemistry and Physics, 18, 17769-17800, https://doi.org/10.5194/acp-18-
17769-2018, 2018.
995
Nault, B. A., Jo, D. S., McDonald, B. C., Campuzano-Jost, P., Day, D. A., Hu, W., Schroder, J. C., Allan, J., Blake, D. R.,
Canagaratna, M. R., Coe, H., Coggon, M. M., DeCarlo, P. F., Diskin, G. S., Dunmore, R., Flocke, F., Fried, A., Gilman,
J. B., Gkatzelis, G., Hamilton, J. F., Hanisco, T. F., Hayes, P. L., Henze, D. K., Hodzic, A., Hopkins, J., Hu, M., Huey,
L. G., Jobson, B. T., Kuster, W. C., Lewis, A., Li, M., Liao, J., Nawaz, M. O., Pollack, I. B., Peischl, J., Rappengck,
B., Reeves, C. E., Richter, D., Roberts, J. M., Ryerson, T. B., Shao, M., Sommers, J. M., Walega, J., Warneke, C.,
1000
Weibring, P., Wolfe, G. M., Young, D. E., Yuan, B., Zhang, Q., de Gouw, J. A., and Jimenez, J. L.: Secondary organic
aerosols from anthropogenic volatile organic compounds contribute substantially to air pollution mortality,
Atmospheric Chemistry and Physics, 21, 11201-11224, https://doi.org/10.5194/acp-21-11201-2021, 2021.
Nel, A.: Atmosphere. Air pollution-related illness: effects of particles, Science, 308, 804-806,
https://doi.org/10.1126/science.1108752, 2005.
1005
Ni, J., Yu, G., Guo, Q., Liang, Q., and Zhou, Z.: Experimental and Numerical Study of the Flow Field and Temperature
Field for a Large-Scale Radiant Syngas Cooler, Industrial & Engineering Chemistry Research, 49, 4452-4461,
https://doi.org/10.1021/ie100014r, 2010.
Ortega, A. M., Day, D. A., Cubison, M. J., Brune, W. H., Bon, D., de Gouw, J. A., and Jimenez, J. L.: Secondary organic
aerosol formation and primary organic aerosol oxidation from biomass-burning smoke in a flow reactor during
1010
FLAME-3, Atmospheric Chemistry and Physics, 13, 11551-11571, https://doi.org/10.5194/acp-13-11551-2013, 2013.
Ortega, A. M., Hayes, P. L., Peng, Z., Palm, B. B., Hu, W., Day, D. A., Li, R., Cubison, M. J., Brune, W. H., Graus, M.,
Warneke, C., Gilman, J. B., Kuster, W. C., de Gouw, J., Gutiérrez-Montes, C., and Jimenez, J. L.: Real-time
measurements of secondary organic aerosol formation and aging from ambient air in an oxidation flow reactor in the
Los Angeles area, Atmospheric Chemistry and Physics, 16, 7411-7433, https://doi.org/10.5194/acp-16-7411-2016,
1015
2016.
Paciga, A., Karnezi, E., Kostenidou, E., Hildebrandt, L., Psichoudaki, M., Engelhart, G. J., Lee, B. H., Crippa, M., Prévôt,
A. S. H., Baltensperger, U., and Pandis, S. N.: Volatility of organic aerosol and its components in the megacity of
Paris, Atmos. Chem. Phys., 16, 2013-2023, https://doi.org/10.5194/acp-16-2013-2016, 2016.
Palm, B. B., Campuzano-Jost, P., Ortega, A. M., Day, D. A., Kaser, L., Jud, W., Karl, T., Hansel, A., Hunter, J. F., Cross,
1020
E. S., Kroll, J. H., Peng, Z., Brune, W. H., and Jimenez, J. L.: In situ secondary organic aerosol formation from ambient
pine forest air using an oxidation flow reactor, Atmospheric Chemistry and Physics, 16, 2943-2970,
https://doi.org/10.5194/acp-16-2943-2016, 2016.
Palm, B. B., Campuzano-Jost, P., Day, D. A., Ortega, A. M., Fry, J. L., Brown, S. S., Zarzana, K. J., Dube, W., Wagner,
N. L., Draper, D. C., Kaser, L., Jud, W., Karl, T., Hansel, A., Gutiérrez-Montes, C., and Jimenez, J. L.: Secondary
1025
organic aerosol formation from in situ OH, O3, and NO3 oxidation of ambient forest air in an oxidation flow reactor,
Atmospheric Chemistry and Physics, 17, 5331-5354, https://doi.org/10.5194/acp-17-5331-2017, 2017.
Palm, B. B., de , S. S., Day, D. A., Campuzano-Jost, P., Hu, W., Seco, R., Sjostedt, S. J., Park, J.-H., Guenther, A. B.,
Kim, S., Brito, J., Wurm, F., Artaxo, P., Thalman, R., Wang, J., Yee, L. D., Wernis, R., Isaacman -VanWertz, G.,
Goldstein, A. H., Liu, Y., Springston, S. R., Souza, R., Newburn, M. K., Alexander, M. L., Martin, S. T., and Jimenez,
1030
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
31
J. L.: Secondary organic aerosol formation from ambient air in an oxidation flow reactor in central Amazonia,
Atmospheric Chemistry and Physics, 18, 467-493, https://doi.org/10.5194/acp-18-467-2018, 2018.
Pankow, J. F.: An absorption model of gas/particle partitioning of organic compounds in the atmosphere, Atmospheric
Environment, 28, 185-188, https://doi.org/10.1016/1352-2310(94)90093-0, 1994.
Park, G., Kim, K., Kang, S., Park, T., Ban, J., and Lee, T.: The Chemical Characteristics and Formation of Potential
1035
Secondary Aerosol (PSA) using an Oxidation Flow Reactor (OFR) in the Summer: Focus on the Residential Area,
Suwon, Journal of Korean Society for Atmospheric Environment, 35, 786-801,
https://doi.org/10.5572/kosae.2019.35.6.786, 2019.
Park, Y. and Kim, H.: Real time measurements of the secondary organic aerosol formation and aging from ambient air
using an oxidation flow reactor in seoul during winter, Environ Pollut, 327, 121464,
1040
https://doi.org/10.1016/j.envpol.2023.121464, 2023.
Patankar, S. V. and Spalding, D. B.: A calculation procedure for heat, mass and momentum transfer in three-dimensional
parabolic flows, International Journal of Heat and Mass Transfer, 15, 1787-1806, https://doi.org/10.1016/0017-
9310(72)90054-3, 1972.
Pathak, R. K., Stanier, C. O., Donahue, N. M., and Pandis, S. N.: Ozonolysis of α-pinene at atmospherically relevant
1045
concentrations: Temperature dependence of aerosol mass fractions (yields), Journal of Geophysical Research, 112,
D03201, https://doi.org/10.1029/2006jd007436, 2007.
Paulsen, D., Dommen, J., Kalberer, M., Prevot, A. S., Richter, R., Sax, M., Steinbacher, M., Weingartner, E., and
Baltensperger, U.: Secondary organic aerosol formation by irradiation of 1,3,5-trimethylbenzene-NOx-H2O in a new
reaction chamber for atmospheric chemistry and physics, Environ Sci Technol, 39, 2668-2678,
1050
https://doi.org/10.1021/es0489137, 2005.
Peng, Z. and Jimenez, J. L.: Radical chemistry in oxidation flow reactors for atmospheric chemistry research, Chem Soc
Rev, 49, 2570-2616, https://doi.org/10.1039/c9cs00766k, 2020.
Peng, Z., Lee-Taylor, J., Orlando, J. J., Tyndall, G. S., and Jimenez, J. L.: Organic peroxy radical chemistry in oxidation
flow reactors and environmental chambers and their atmospheric relevance, Atmospheric Chemistry and Physics, 19,
1055
813-834, https://doi.org/10.5194/acp-19-813-2019, 2019.
Peng, Z., Day, D. A., Stark, H., Li, R., Lee-Taylor, J., Palm, B. B., Brune, W. H., and Jimenez, J. L.: HOx radical chemistry
in oxidation flow reactors with low-pressure mercury lamps systematically examined by modeling, Atmospheric
Measurement Techniques, 8, 4863-4890, https://doi.org/10.5194/amt-8-4863-2015, 2015.
Peng, Z., Palm, B. B., Day, D. A., Talukdar, R. K., Hu, W., Lambe, A. T., Brune, W. H., and Jimenez, J. L.: Model
1060
Evaluation of New Techniques for Maintaining High-NO Conditions in Oxidation Flow Reactors for the Study of
OH-Initiated Atmospheric Chemistry, ACS Earth and Space Chemistry, 2, 72-86,
https://doi.org/10.1021/acsearthspacechem.7b00070, 2018.
Pierce, J. R., Chen, K., and Adams, P. J.: Contribution of primary carbonaceous aerosol to cloud condensation nuclei:
processes and uncertainties evaluated with a global aerosol microphysics model, Atmospheric Chemistry and Physics,
1065
7, 5447-5466, https://doi.org/10.5194/acp-7-5447-2007, 2007.
Poschl, U.: Atmospheric aerosols: composition, transformation, climate and health effects, Angew Chem Int Ed Engl, 44,
7520-7540, https://doi.org/10.1002/anie.200501122, 2005.
Presto, A. A., Hartz, K. E., and Donahue, N. M.: Secondary organic aerosol production from terpene ozonolysis. 2. Effect
of NOx concentration, Environ Sci Technol, 39, 7046-7054, https://doi.org/10.1021/es050400s, 2005.
1070
Price, D. J., Kacarab, M., Cocker, D. R., Purvis-Roberts, K. L., and Silva, P. J.: Effects of temperature on the formation of
secondary organic aerosol from amine precursors, Aerosol Science and Technology, 50, 1216-1226,
https://doi.org/10.1080/02786826.2016.1236182, 2016.
Qi, L., Nakao, S., Tang, P., and Cocker, D. R.: Temperature effect on physical and chemical properties of secondary organic
aerosol from m-xylene photooxidation, Atmospheric Chemistry and Physics, 10, 3847-3854,
1075
https://doi.org/10.5194/acp-10-3847-2010, 2010.
Quéver, L. L. J., Kristensen, K., Normann Jensen, L., Rosati, B., Teiwes, R., Daellenbach, K. R., Peräky, O., Roldin,
P., Bossi, R., Pedersen, H. B., Glasius, M., Bilde, M., and Ehn, M.: Effect of temperature on the formation of highly
oxygenated organic molecules (HOMs) from alpha-pinene ozonolysis, Atmospheric Chemistry and Physics, 19, 7609-
7625, https://doi.org/10.5194/acp-19-7609-2019, 2019.
1080
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
32
Riipinen, I., Pierce, J. R., Donahue, N. M., and Pandis, S. N.: Equilibration time scales of organic aerosol inside
thermodenuders: Evaporation kinetics versus thermodynamics, Atmospheric Environment, 44, 597-607,
https://doi.org/10.1016/j.atmosenv.2009.11.022, 2010.
Robbins, R. C. and Cadle, R. D.: Kinetics of the Reaction between Gaseous Ammonia and Sulfuric Acid Droplets in an
Aerosol, The Journal of Physical Chemistry, 62, 469-471, https://doi.org/10.1021/j150562a025, 2002.
1085
Roldin, P., Eriksson, A. C., Nordin, E. Z., Hermansson, E., Mogensen, D., Rusanen, A., Boy, M., Swietlicki, E.,
Svenningsson, B., Zelenyuk, A., and Pagels, J.: Modelling non-equilibrium secondary organic aerosol formation and
evaporation with the aerosol dynamics, gas- and particle-phase chemistry kinetic multilayer model ADCHAM,
Atmospheric Chemistry and Physics, 14, 7953-7993, https://doi.org/10.5194/acp-14-7953-2014, 2014.
Rollins, A. W., Kiendler-Scharr, A., Fry, J. L., Brauers, T., Brown, S. S., Dorn, H. P., Dubé, W. P., Fuchs, H., Mensah, A.,
1090
Mentel, T. F., Rohrer, F., Tillmann, R., Wegener, R., Wooldridge, P. J., and Cohen, R. C.: Isoprene oxidation by
nitrate radical: alkyl nitrate and secondary organic aerosol yields, Atmospheric Chemistry and Physics, 9, 6685-6703,
https://doi.org/10.5194/acp-9-6685-2009, 2009.
Rowe, J. P., Lambe, A. T., and Brune, W. H.: Technical Note: Effect of varying the λ = 185 and 254 nm photon flux ratio
on radical generation in oxidation flow reactors, ACP, https://doi.org/10.5194/acp-2020-642, 2020.
1095
Saha, P. K. and Grieshop, A. P.: Exploring Divergent Volatility Properties from Yield and Thermodenuder Measurements
of Secondary Organic Aerosol from alpha-Pinene Ozonolysis, Environ Sci Technol, 50, 5740-5749,
https://doi.org/10.1021/acs.est.6b00303, 2016.
Saha, P. K., Reece, S. M., and Grieshop, A. P.: Seasonally Varying Secondary Organic Aerosol Formation From In-Situ
Oxidation of Near-Highway Air, Environ Sci Technol, 52, 7192-7202, https://doi.org/10.1021/acs.est.8b01134, 2018.
1100
Saha, P. K., Khlystov, A., Yahya, K., Zhang, Y., Xu, L., Ng, N. L., and Grieshop, A. P.: Quantifying the volatility of
organic aerosol in the southeastern US, Atmospheric Chemistry and Physics, 17, 501-520, https://doi.org/10.5194/acp-
17-501-2017, 2017.
Sbai, S. E., Li, C., Boreave, A., Charbonnel, N., Perrier, S., Vernoux, P., Bentayeb, F., George, C., and Gil, S.: Atmospheric
photochemistry and secondary aerosol formation of urban air in Lyon, France, J Environ Sci (China), 99, 311-323,
1105
https://doi.org/10.1016/j.jes.2020.06.037, 2021.
Schobesberger, S., D'Ambro, E. L., Lopez-Hilfiker, F. D., Mohr, C., and Thornton, J. A.: A model framework to retrieve
thermodynamic and kinetic properties of organic aerosol from composition-resolved thermal desorption
measurements, Atmospheric Chemistry and Physics, 18, 14757-14785, https://doi.org/10.5194/acp-18-14757-2018,
2018.
1110
Sengupta, D., Samburova, V., Bhattarai, C., Kirillova, E., Mazzoleni, L., Iaukea-Lum, M., Watts, A., Moosmüller, H., and
Khlystov, A.: Light absorption by polar and non-polar aerosol compounds from laboratory biomass combustion,
Atmospheric Chemistry and Physics, 18, 10849-10867, https://doi.org/10.5194/acp-18-10849-2018, 2018.
Shah, R. U., Coggon, M. M., Gkatzelis, G. I., McDonald, B. C., Tasoglou, A., Huber, H., Gilman, J., Warneke, C., Robinson,
A. L., and Presto, A. A.: Urban Oxidation Flow Reactor Measurements Reveal Significant Secondary Organic Aerosol
1115
Contributions from Volatile Emissions of Emerging Importance, Environ Sci Technol, 54, 714-725,
https://doi.org/10.1021/acs.est.9b06531, 2020.
Shih, T.-H., Liou, W. W., Shabbir, A., Yang, Z., and Zhu, J.: A new k-ϵ eddy viscosity model for high reynolds number
turbulent flows, Computers & Fluids, 24, 227-238, https://doi.org/10.1016/0045-7930(94)00032-t, 1995.
Siemens, K., Morales, A., He, Q., Li, C., Hettiyadura, A. P. S., Rudich, Y., and Laskin, A.: Molecular Analysis of
1120
Secondary Brown Carbon Produced from the Photooxidation of Naphthalene, Environ Sci Technol, 56, 3340-3353,
https://doi.org/10.1021/acs.est.1c03135, 2022.
Simon, M., Dada, L., Heinritzi, M., Scholz, W., Stolzenburg, D., Fischer, L., Wagner, A. C., rten, A., Rörup, B., He,
X.-C., Almeida, J., Baalbaki, R., Baccarini, A., Bauer, P. S., Beck, L., Bergen, A., Bianchi, F., Bräkling, S., Brilke,
S., Caudillo, L., Chen, D., Chu, B., Dias, A., Draper, D. C., Duplissy, J., El-Haddad, I., Finkenzeller, H., Frege, C.,
1125
Gonzalez-Carracedo, L., Gordon, H., Granzin, M., Hakala, J., Hofbauer, V., Hoyle, C. R., Kim, C., Kong, W.,
Lamkaddam, H., Lee, C. P., Lehtipalo, K., Leiminger, M., Mai, H., Manninen, H. E., Marie, G., Marten, R., Mentler,
B., Molteni, U., Nichman, L., Nie, W., Ojdanic, A., Onnela, A., Partoll, E., Petä, T., Pfeifer, J., Philippov, M.,
Quéver, L. L. J., Ranjithkumar, A., Rissanen, M. P., Schallhart, S., Schobesberger, S., Schuchmann, S., Shen, J.,
Sipilä, M., Steiner, G., Stozhkov, Y., Tauber, C., Tham, Y. J., Tomé, A. R., Vazquez-Pufleau, M., Vogel, A. L.,
1130
Wagner, R., Wang, M., Wang, D. S., Wang, Y., Weber, S. K., Wu, Y., Xiao, M., Yan, C., Ye, P., Ye, Q., Zauner-
Wieczorek, M., Zhou, X., Baltensperger, U., Dommen, J., Flagan, R. C., Hansel, A., Kulmala, M., Volkamer, R.,
Winkler, P. M., Worsnop, D. R., Donahue, N. M., Kirkby, J., and Curtius, J.: Molecular understanding of new-particle
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
33
formation from α-pinene between −50 and +25 C, Atmospheric Chemistry and Physics, 20, 9183-9207,
https://doi.org/10.5194/acp-20-9183-2020, 2020.
1135
Simonen, P., Saukko, E., Karjalainen, P., Timonen, H., Bloss, M., Aakko-Saksa, P., Rönkkö, T., Keskinen, J., and Dal
Maso, M.: A new oxidation flow reactor for measuring secondary aerosol formation of rapidly changing emission
sources, Atmospheric Measurement Techniques, 10, 1519-1537, https://doi.org/10.5194/amt-10-1519-2017, 2017.
Srivastava, D., Li, W., Tong, S., Shi, Z., and Harrison, R. M.: Characterization of products formed from the oxidation of
toluene and m-xylene with varying NO(x) and OH exposure, Chemosphere, 334, 139002,
1140
https://doi.org/10.1016/j.chemosphere.2023.139002, 2023.
Srivastava, D., Vu, T. V., Tong, S., Shi, Z., and Harrison, R. M.: Formation of secondary organic aerosols from
anthropogenic precursors in laboratory studies, npj Climate and Atmospheric Science, 5, 22,
https://doi.org/10.1038/s41612-022-00238-6, 2022.
Stanier, C. O., Donahue, N., and Pandis, S. N.: Parameterization of secondary organic aerosol mass fractions from smog
1145
chamber data, Atmospheric Environment, 42, 2276-2299, https://doi.org/10.1016/j.atmosenv.2007.12.042, 2008.
Sumlin, B., Fortner, E., Lambe, A., Shetty, N. J., Daube, C., Liu, P., Majluf, F., Herndon, S., and Chakrabarty, R. K.: Diel
cycle impacts on the chemical and light absorption properties of organic carbon aerosol from wildfires in the western
United States, Atmospheric Chemistry and Physics, 21, 11843-11856, https://doi.org/10.5194/acp-21-11843-2021,
2021.
1150
Takekawa, H., Minoura, H., and Yamazaki, S.: Temperature dependence of secondary organic aerosol formation by photo-
oxidation of hydrocarbons, Atmospheric Environment, 37, 3413-3424, https://doi.org/10.1016/s1352-2310(03)00359-
5, 2003.
Tillmann, R., Hallquist, M., Jonsson, Å. M., Kiendler-Scharr, A., Saathoff, H., Iinuma, Y., and Mentel, T. F.: Influence of
relative humidity and temperature on the production of pinonaldehyde and OH radicals from the ozonolysis of α -
1155
pinene, Atmospheric Chemistry and Physics, 10, 7057-7072, https://doi.org/10.5194/acp-10-7057-2010, 2010.
Tkacik, D. S., Lambe, A. T., Jathar, S., Li, X., Presto, A. A., Zhao, Y., Blake, D., Meinardi, S., Jayne, J. T., Croteau, P. L.,
and Robinson, A. L.: Secondary organic aerosol formation from in-use motor vehicle emissions using a potential
aerosol mass reactor, Environ Sci Technol, 48, 11235-11242, https://doi.org/10.1021/es502239v, 2014.
Wang, X., Liu, T., Bernard, F., Ding, X., Wen, S., Zhang, Y., Zhang, Z., He, Q., Lü, S., Chen, J., Saunders, S., and Yu, J.:
1160
Design and characterization of a smog chamber for studying gas-phase chemical mechanisms and aerosol formation,
Atmospheric Measurement Techniques, 7, 301-313, https://doi.org/10.5194/amt-7-301-2014, 2014.
Warren, B., Austin, R. L., and Cocker, D. R.: Temperature dependence of secondary organic aerosol, Atmospheric
Environment, 43, 3548-3555, https://doi.org/10.1016/j.atmosenv.2009.04.011, 2009.
Watne, A. K., Psichoudaki, M., Ljungstrom, E., Le Breton, M., Hallquist, M., Jerksjo, M., Fallgren, H., Jutterstrom, S.,
1165
and Hallquist, A. M.: Fresh and Oxidized Emissions from In-Use Transit Buses Running on Diesel, Biodiesel, and
CNG, Environ Sci Technol, 52, 7720-7728, https://doi.org/10.1021/acs.est.8b01394, 2018.
White, S., Angove, D., Li, K., Campbell, I., Element, A., Halliburton, B., Lavrencic, S., Cameron, D., Jamie, I., and Azzi,
M.: Development of a new smog chamber for studying the impact of different UV lamps on SAPRC chemical
mechanism predictions and aerosol formation, Environmental Chemistry, 15, https://doi.org/10.1071/en18005, 2018.
1170
Xu, N. and Collins, D. R.: Design and characterization of a new oxidation flow reactor for laboratory and long-term ambient
studies, Atmospheric Measurement Techniques, 14, 2891-2906, https://doi.org/10.5194/amt-14-2891-2021, 2021.
Xu, W., Li, Z., Lambe, A. T., Li, J., Liu, T., Du, A., Zhang, Z., Zhou, W., and Sun, Y.: Secondary organic aerosol formation
and aging from ambient air in an oxidation flow reactor during wintertime in Beijing, China, Environ Res, 209, 112751,
https://doi.org/10.1016/j.envres.2022.112751, 2022.
1175
Xu, W., Takeuchi, M., Chen, C., Qiu, Y., Xie, C., Xu, W., Ma, N., Worsnop, D. R., Ng, N. L., and Sun, Y.: Estimation of
particulate organic nitrates from thermodenuder-aerosol mass spectrometer measurements in North China Plain,
Atmospheric Measurement Techniques Discussions, https://doi.org/10.5194/amt-2020-474, 2020.
Zhang, X., McVay, R. C., Huang, D. D., Dalleska, N. F., Aumont, B., Flagan, R. C., and Seinfeld, J. H.: Formation and
evolution of molecular products in alpha-pinene secondary organic aerosol, Proc Natl Acad Sci U S A, 112, 14168-
1180
14173, https://doi.org/10.1073/pnas.1517742112, 2015.
Zhao, R., Zhang, Q., Xu, X., Zhao, W., Yu, H., Wang, W., Zhang, Y., and Zhang, W.: Effect of experimental conditions
on secondary organic aerosol formation in an oxidation flow reactor, Atmospheric Pollution Research, 12, 205-213,
https://doi.org/10.1016/j.apr.2021.01.011, 2021.
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
34
Ziemann, P. J. and Atkinson, R.: Kinetics, products, and mechanisms of secondary organic aerosol formation, Chem Soc
1185
Rev, 41, 6582-6605, https://doi.org/10.1039/c2cs35122f, 2012.
Zong, T., Wu, Z., Wang, J., Bi, K., Fang, W., Yang, Y., Yu, X., Bao, Z., Meng, X., Zhang, Y., Guo, S., Chen, Y., Liu, C.,
Zhang, Y., Li, S. M., and Hu, M.: A new smog chamber system for atmospheric multiphase chemistry study: design
and characterization, Atmos. Meas. Tech. Discuss., 2023, 1-25, https://doi.org/10.5194/amt-2023-34, 2023.
1190
https://doi.org/10.5194/amt-2023-230
Preprint. Discussion started: 16 November 2023
c
Author(s) 2023. CC BY 4.0 License.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Aromatic volatile organic compounds (VOCs) are an important precursor of secondary organic aerosol (SOA) in the urban environment. SOA formed from the oxidation of anthropogenic VOCs can be substantially more abundant than biogenic SOA and has been shown to account for a significant fraction of fine particulate matter in urban areas. A potential aerosol mass (PAM) chamber was used to investigate the oxidised products from the photo-oxidation of m-xylene and toluene. The experiments were carried out with OH radical as oxidant in both high- and low-NOx conditions and the resultant aerosol samples were collected using quartz filters and analysed by GC × GC-TOFMS. Results show the oxidation products derived from both precursors included ring-retaining and -opening compounds (unsaturated aldehydes, unsaturated ketones and organic acids) with a high number of ring-opening compounds observed from toluene oxidation. Glyoxal and methyl glyoxal were the major ring-cleavage products from both oxidation systems, indicating that a bicyclic route plays an important role in their formation. SOA yields were higher for both precursors under high-NOx (toluene: 0.111; m-xylene: 0.124) than at low-NOx (toluene: 0.089; m-xylene: 0.052), likely linked to higher OH concentrations during low-NOx experiments which may lead to higher degree of fragmentation. DHOPA (2,3-dihydroxy-4-oxo-pentanoic acid), a known tracer of toluene oxidation, was observed in both oxidation systems. The mass fraction of DHOPA in SOA from toluene oxidation was about double the value reported previously, but it should not be regarded as a tracer solely for oxidation of toluene as m-xylene oxidation gave a similar relative yield.
Article
Full-text available
To investigate the impact of aging on the sources, volatility, and viscosity of organic aerosol (OA) in Chinese outflows, a high-resolution time-of-flight aerosol mass spectrometer (HR-AMS) coupled with a thermodenuder (TD) was deployed in the spring of 2018 in Dongying, which is a regional receptor site of metropolitan emissions in the North China Plain (NCP). The average mass concentration of PM1 is 31.5±22.7 µgm-3, which is mainly composed of nitrate (33 %) and OA (25 %). The source apportionment results show that the OA is mainly contributed by oxygenated OA (OOA) from secondary sources, including background-OOA (33 %) representing a background concentration of OA (2.6 µgm-3) in the NCP area, and transported-OOA (33 %) oxidized from urban emissions. The other two factors include aged hydrocarbon-liked OA (aged-HOA, 28 %) from transported vehicle emissions and biomass burning OA (BBOA, 5 %) from local open burning. The volatility of total OA (average C*=3.2×10-4 µgm-3) in this study is generally lower than that reported in previous field studies, which is mainly due to the high OA oxidation level resulting from aging processes during transport. The volatilities of OA factors follow the order of background-OOA (average C*=2.7×10-5 µgm-3) < transported-OOA (3.7×10-4µgm-3)< aged-HOA (8.1×10-4µgm-3)< BBOA (0.012 µgm-3). Extremely low volatilities in ambient air indicate that oligomers may exist in aged plumes. The viscosity estimation suggests that the majority of ambient OA in this study behaves as semisolid (60 %), liquifies at higher relative humidity (RH) (21 %), and solidifies (19 %) during noon when the RH is low and the oxidation level is high. Finally, the estimated mixing time of molecules in 200 nm OA varies dramatically from minutes at night to years in the afternoon, emphasizing the need to consider its dynamic kinetic limits when modeling OA. In general, the overall results of this study improve our understanding of the impact of aging on OA volatility and viscosity.
Article
Full-text available
2022): The organic coating unit, an all-in-one system for reproducible generation of secondary organic matter aerosol, Aerosol Science and Technology, ABSTRACT We report on a novel automated oxidation flow reactor to generate a wide variety of organic aerosol samples. The instrument is equipped with a humidifier, a dosing system for volatile organic precursors and an oxidation flow reactor (OFR) for generation of secondary organic matter (SOM). The instrument, known as organic coating unit (OCU), can produce homogeneously nucleated SOM particles or, used in combination with a standard combustion generator (e.g., a diffusion flame soot generator or any other seed particle), particles coated with a controlled amount of SOM. The physical and chemical properties of the generated particles can be controlled in a simple manner by selecting through a touch-screen target values for parameters, such as organic gaseous precursor concentration, humidity, and UV (ultraviolet) light intensity. Parameters and measured quantities are automatically stored in text files for easy export and analysis. Furthermore, we provide stable operation conditions and characterize the physicochemical properties of the generated aerosols with an array of methods, including transmission electron microscopy (TEM), thermal-optical analysis and liquid chromatography coupled with mass spectrometry (LC-MS). This all-in-one instrument is robust, compact, portable, and user-friendly, making it ideal for laboratory or field-based aerosol studies. ARTICLE HISTORY
Article
Full-text available
caryophyllene (BCP) is one of the most important sesquiterpenes (SQTs) in the atmosphere, with a large potential contribution to secondary organic aerosol (SOA) formation mainly from reactions with ozone (O3) and nitrate radicals (NO3). In this work, we study the temperature dependence of the kinetics of BCP ozonolysis, SOA yields, and SOA chemical composition in the dark and in the absence and presence of nitrogen oxides including nitrate radicals (NO3). We cover a temperature range of 213–313 K, representative of tropospheric conditions. The oxidized components in both gas and particle phases were characterized on a molecular level by a chemical ionization mass spectrometer equipped with a filter inlet for gases and aerosols using iodide as the reagent ion (FIGAERO-iodide-CIMS). The batch mode experiments were conducted in the 84.5 m3 aluminium simulation chamber AIDA at the Karlsruhe Institute of Technology (KIT). In the absence of nitrogen oxides, the temperature-dependent rate coefficient of the endocyclic double bond in BCP reacting with ozone between 243–313 K is negatively correlated with temperature, corresponding to the following Arrhenius equation: k= (1.6 ± 0.4) × 10-15 × exp((559 ± 97)/T). The SOA yields increase from 16 ± 5 % to 37 ± 11 %, with temperatures decreasing from 313 to 243 K at a total organic particle mass of 10 µg m-3. The variation in the ozonolysis temperature leads to a substantial impact on the abundance of individual organic molecules. In the absence of nitrogen oxides, monomers C14-15H22-24O3-7 (37.4 %), dimers C28-30H44-48O5-9 (53.7 %), and trimers C41-44H62-66O9-11 (8.6 %) are abundant in the particle phase at 213 K. At 313 K, we observed more oxidized monomers (mainly C14-15H22-24O6-9, 67.5 %) and dimers (mainly C27-29H42-44O9-11, 27.6 %), including highly oxidized molecules (HOMs; C14H22O7,9, C15H22O7,9C15H24O7,9), which can be formed via hydrogen shift mechanisms, but no significant trimers. In the presence of nitrogen oxides, the organonitrate fraction increased from 3 % at 213 K to 12 % and 49 % at 243 and 313 K, respectively. Most of the organonitrates were monomers with C15 skeletons and only one nitrate group. More highly oxygenated organonitrates were observed at higher temperatures, with their signal-weighted O:C atomic ratio increasing from 0.41 to 0.51 from 213 to 313 K. New dimeric and trimeric organic species without nitrogen atoms (C20, C35) were formed in the presence of nitrogen oxides at 298–313 K, indicating potential new reaction pathways. Overall, our results show that increasing temperatures lead to a relatively small decrease in the rate coefficient of the endocyclic double bond in BCP reacting with ozone but to a strong decrease in SOA yields. In contrast, the formation of HOMs and organonitrates increases significantly with temperature.
Article
Full-text available
The chemical pathways for the production of secondary organic aerosols (SOA) are influenced by the concentration of nitrogen oxides (NOx), including the production of organonitrates (ON). Herein, a series of experiments conducted in an environmental chamber investigated the production and partitioning of total organonitrates from α-pinene photo-oxidation from <1 to 24 ppb NOx. Gas-phase and particle-phase organonitrates (gON and pON, respectively) were measured by laser-induced fluorescence (LIF). The composition of the particle phase and the particle mass concentration were simultaneously characterized by online aerosol mass spectrometry. The LIF and MS measurements of pON concentrations had a Pearson correlation coefficient of 0.91 from 0.3 to 1.1 μg m-3. For 1-6 ppb NOx, the yield of SOA particle mass concentration increased from 0.02 to 0.044 with NOx concentration. For >6 ppb NOx, the yield steadily dropped, reaching 0.034 at 24 ppb NOx. By comparison, the yield of pON steadily increased from 0.002 to 0.022 across the range of investigated NOx concentrations. The yield of gON likewise increased from 0.005 to 0.148. The gas-to-particle partitioning ratio (pON/(pON + gON)) depended strongly on the NOx concentration, changing from 0.27 to 0.13 as the NOx increased from <1 to 24 ppb. In the atmosphere, there is typically a cross-over point between clean and polluted conditions that strongly affects SOA production, and the results herein quantitatively identify 6 ppb NOx as that point for α-pinene photo-oxidation under these study conditions, including the production and partitioning of organonitrates. The trends in SOA yield and partitioning ratio as a function of NOx occur because of the changes in pON volatility.
Article
Full-text available
Aerosols produced from the oxidation of volatile/semi-volatile organic compounds (VOCs/SVOCs), known as secondary organic aerosol (SOA), account for a significant fraction of atmospheric airborne particles. This paper reviews the current understanding of SOA formation from gas-phase oxidation with a focus on anthropogenic precursors and their reaction products from atmospheric simulation chamber studies. The review summarises the major reaction products derived from main groups of SOA precursors (e.g., alkanes, aromatics), SOA yields and the factors controlling SOA formation. We highlight that lab-derived SOA yield depends strongly upon, not only the concentrations of SOA precursors and oxidants but also simulation conditions. npj Climate and Atmospheric Science (2022) 5:22 ; https://doi.
Article
Herein, the formation and aging processes of organic aerosol (OA) in urban Seoul, Korea, during winter were investigated using a high-resolution aerosol mass spectrometer (HR-ToF-AMS) and an oxidation flow reactor (OFR). The results demonstrated that the highest secondary OA (SOA) production (ΔOA = 3.44 μg m-3 with a relative OA enhancement ratio (EROA) = 1.40) occurred at ∼2 eq. days of OH exposure. Particularly, higher SOA production was observed under the following atmospheric conditions: high relative humidity (RH) (>70%) and high PM1 mass concentration (>50 μg m-3), demonstrating that oxidation capacity, heterogeneous and aqueous phase reactions are important for further oxidation. Additionally, increased SOA formation occurs under both higher hydrocarbon-like OA and more oxidized OOA conditions. Further oxidation of both freshly emitted and aged and/or transported OA can be a remarkable further source of SOA in winter in Seoul and further downwind areas. In particular, the high mass concentration of MO-OOA in high total PM1 would be an important indication that SOA formation could be accelerated by a heterogeneous reaction, necessitating additional investigations on the haze formation process.
Article
Secondary organic aerosol (SOA) data gathered in environmental chambers (ECs) have been used extensively to develop parameters to represent SOA formation and evolution. The EC-based parameters are usually constrained to less than one day of photochemical aging but extrapolated to predict SOA aging over much longer timescales in atmospheric models. Recently, SOA has been increasingly studied in oxidation flow reactors (OFRs) over aging timescales of one to multiple days. However, these OFR data have been rarely used to validate or update the EC-based parameters. The simultaneous use of EC and OFR data is challenging because the processes relevant to SOA formation and evolution proceed over very different timescales, and both reactor types exhibit distinct experimental artifacts. In this work, we show that a kinetic SOA chemistry and microphysics model that accounts for various processes, including wall losses, aerosol phase state, heterogeneous oxidation, oligomerization, and new particle formation, can simultaneously explain SOA evolution in EC and OFR experiments, using a single consistent set of SOA parameters. With α-pinene as an example, we first developed parameters by fitting the model output to the measured SOA mass concentration and oxygen-to-carbon (O:C) ratio from an EC experiment (<1 day of aging). We then used these parameters to simulate SOA formation in OFR experiments and found that the model overestimated SOA formation (by a factor of 3-16) over photochemical ages ranging from 0.4 to 13 days, when excluding the abovementioned processes. By comprehensively accounting for these processes, the model was able to explain the observed evolution in SOA mass, composition (i.e., O:C), and size distribution in the OFR experiments. This work suggests that EC and OFR SOA data can be modeled consistently, and a synergistic use of EC and OFR data can aid in developing more refined SOA parameters for use in atmospheric models.