Next Article in Journal
Analysis of Separation Distances under Varying Odour Emission Rates and Meteorology: A WWTP Case Study
Next Article in Special Issue
The Ångström Exponent and Single-Scattering Albedo of Black Carbon: Effects of Different Coating Materials
Previous Article in Journal
The Spatio—Temporal Variation of Pacific Blocking Frequency within Winter Months and Its Relationship with Surface Air Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Ammonium Sulfate Seed Particle on Optics and Compositions of Toluene Derived Organic Aerosol in Photochemistry

1
Fujian Province Key Laboratory of Modern Analytical Science and Separation Technology, College of Chemistry & Chemical Engineering and Environment, Minnan Normal University, Zhangzhou 363000, China
2
Laboratory of Atmospheric Physico-Chemistry, Anhui Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Hefei 230031, China
*
Authors to whom correspondence should be addressed.
Atmosphere 2020, 11(9), 961; https://doi.org/10.3390/atmos11090961
Submission received: 15 July 2020 / Revised: 6 August 2020 / Accepted: 8 August 2020 / Published: 10 September 2020
(This article belongs to the Special Issue Carbonaceous Aerosol)

Abstract

:
Aromatic secondary organic aerosol (SOA) particles are known to contribute to radiative forcing and light absorption of atmosphere. However, the complex refractive index (CRI), single-scattering albedo (SSA) and other optical parameters of aromatic SOA are not well understood. SOA generated from photooxidation of toluene with a variety concentration of ammonium sulfate ((NH4)2SO4) seed particles in a smog chamber were investigated in the current study. The real part CRI of toluene SOA without seeds derived and based on aerosol albedometer measurements is 1.486 ± 0.002 at λ = 470 nm, showing a good agreement with available experimental data, and its SSA was measured to be 0.92 ± 0.02 at λ = 470 nm, indicating that the SOA particles without seeds have strong scattering ability. The SSA of SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed was 0.81 ± 0.02 at λ = 470 nm, less than the SSA of SOA without seed. SSA of SOA decreased, while the imaginary part of CRI (k) of SOA increased with increasing concentration of (NH4)2SO4 seed, demonstrating that the adsorption capacity of SOA formed in the presence of (NH4)2SO4 seed is enhanced. Different from the carboxyl compounds measured in the SOA without seed, imidazoles with strong chromophores of C=N that are responsible for the light absorption were detected as the principal constituents of SOA formed in the presence of (NH4)2SO4 seed. These would provide valuable information for discussing the optics and components of aromatic SOA in the urban atmosphere containing a high concentration of (NH4)2SO4 fine particles.

1. Introduction

The photooxidation of toluene and other monocyclic aromatic compounds from anthropogenic emission sources leads to the formation of semi- and non-volatile oxygenated compounds, which result in secondary organic aerosol (SOA) particles [1,2]. Interest in SOA has been renewed because of its contribution to radiative balance and visibility degradation [3,4,5]. Generally, the complex refractive index (CRI) and single-scattering albedo (SSA) are used to characterize the fundamental optical parameters of aerosol particles. CRI is expressed as m = n + i k, n and k is the real and imaginary part, corresponding to the scattering and absorption of aerosol particles, respectively. SSA is the ratio of aerosol scattering (αscat) to extinction (αext, the sum of scattering and absorption) coefficient [6]. However, the CRI and SSA of aromatic SOA are still poorly characterized, due to the inherent complexity in aerosol composition and difficulty in accurate measurement of particulate intrinsic optics.
Cavity ring down spectroscopy (CRDS) has been employed in several chamber studies for the investigation of the CRI of aromatic SOA particles generated in the presence of NOx. For toluene SOA particles, Nakayama et al. [7,8] determined the n value of SOA particles to be 1.483 at 532 nm, while the imaginary part of CRI, k-value was observed to increase with increasing NOx concentration. Li et al. [9,10] measured that the real part of CRI of aromatic SOA particles at 532 nm fall in 1.38–1.59, and attributed the decrease of real CRI to the gas-phase partitioning. The real CRI of m-xylene SOA formed without seed and in presence of (NH4)2SO4 seed decreased as time gone on, when CRI was nearly stable, the real part value of CRI for m-xylene SOA was found to decrease by 0.09 and 0.15 compared to the earliest formed SOA, respectively [10]. However, CRDS applied in the work by Nakayama et al. [7,8] and Li et al. [9,10] operated at 532 nm, and the gaseous absorption may bias the extinction result [11]. Additionally, aerosol SSA is not measured in these experiments. Fortunately, the aerosol albedometer can measure the real-time particle optical parameters, such as direct SSA and retrieved CRI [6]. The albedometer generally includes an integrating sphere and incoherent broadband cavity-enhanced absorption spectroscopy (IBBCEAS) for detecting αscat and αext; CRI and SSA can be acquired simultaneously. IBBCEAS can quantitatively measure gaseous absorption and aerosol extinction; gaseous absorption was subtracted for accurate measurement of αext coefficient [12,13]. Recently, our group has developed an albedometer based on IBBCEAS combined with integrating sphere, and measured the SSA of polystyrene latex, ammonium sulfate particles, the optical properties of fine particles and columnar aerosol in Beijing successfully [14,15,16].
The ubiquitous inorganic fine particles contribute substantially to SOA transformation by providing condensation surface in gas-particle partitioning, modifying reaction environment, involving reactions, or promoting their deposition [17,18]. Atmospheric fine particle pollution is serious in the urban areas of China. The average mass concentrations of particles are in the range of 100–300 μg m−3 during haze days in mega-cities of China [19,20]. Ammonium sulfate ((NH4)2SO4) is the major constituent of fine particles during haze days [21]. (NH4)2SO4 fine particles can act as efficient seeds to promote SOA formation, mainly due to their large specific surface area. Moreover, the presence of seed particles shall change the optical and chemical results of SOA. Although the real part of CRI of aromatic SOA have been measured [7,8,9,10], and the real CRI of m-xylene SOA formed in presence of (NH4)2SO4 seed was observed to decrease with reaction time [10], no investigations on the influences of (NH4)2SO4 seed on the SSA of SOA are performed. Thus, experiments were carried out to measure the SSA of toluene SOA formed without, and in the presence of, (NH4)2SO4 seed in the current study. Moreover, αscat and αext coefficients were detected by the aerosol albedometer in real-time, and SSA (=αscat/αext) of the toluene SOA particles were obtained immediately. Additionally, the optical and chemical properties of SOA were further characterized by ultraviolet-visible (UV-Vis) spectrometer and mass spectrometer. The effects of (NH4)2SO4 seed on the optical parameters and the compositions of SOA are discussed in detail.

2. Experiments

2.1. Material

Toluene (>99%) and H2O2 (30%) were supplied by Sigma-Aldrich Chemistry Corporation, Germany. Ammonium sulfate (99.9%) and methanol (>99%) were obtained from The Third Reagent Factory of Tianjin.
Seed particles were generated by aspirating the 4 g/L (NH4)2SO4 solution via the atomizer (TSI Inc, Shoreview, MN, USA, Model 3076), passed through the dryer (TSI Inc, Shoreview, MN, USA, Model 3062) and a neutralizer (TSI Inc, Shoreview, MN, USA, Model 3054) successively, then introduced into the chamber [22,23]. The mean diameter of the (NH4)2SO4 seed was ~100 nm, and seed was established at a certain concentration by controlling the filling time of the seed particles into the chamber.

2.2. Toluene SOA Particles Formation without and in Presence of (NH4)2SO4 Seed

Formation of toluene SOA particles without and in the presence of different concentration of (NH4)2SO4 seed was performed in a 850 L chamber [22,23]. After flushing the chamber, toluene, H2O2 and (NH4)2SO4 seed were delivered into the chamber, which was then filled with zero air to full volume. The concentration of toluene and H2O2 in the chamber was fixed at 600 and 3000 ppb, respectively. Eight experiments were performed without and in presence of (NH4)2SO4 seed in the concentration of 25, 50, 100, 150, 200, 250 and 300 μg/m3, respectively. The temperature in the chamber was kept at about 300 ± 2 K, while relative humidity (RH) was 25 ± 2% for all the experiments. Two small fans were installed inside the flange of the chamber to mix the gases, which ensures the uniform mixing of chemical compounds in the chamber. Four UV lamps were turned on to irradiate H2O2 to generate OH radicals [24], which initiated the photooxidation of toluene to form SOA. As illustrated in Figure 1, the concentrations of NOx, toluene, the mass concentration and optics of SOA particles in the chamber was detected using NO-NO2-NOx analyzer (Thermo Fisher Scientific Inc, Walsham, MA, USA, Model 42i), GC-FID (Agilent Technologies Inc, Palo Alto, CA, USA, Model 7820A), scanning mobility particle sizer (SMPS, TSI Inc, Shoreview, MN, USA, Model 3080L Differential Mobility Analyzer (DMA), Model 3775 Condensation Particle Counter (CPC)) and aerosol albedometer, respectively. Each experiment was conducted for three times, and the average of the results was used as the final data. It is worth noting that no NOx was added for better investigating the effect of (NH4)2SO4 seed aerosol on the optical parameters of toluene SOA. The NOx concentration measured by NO-NO2-NOx analyzer is less than 1 ppb in the whole reaction process. Thus, the NOx concentration in the chamber is not analyzed in this study.

2.3. Aerosol Albedometer

As displayed in Figure 1, light (190–2100 nm) emitted from a laser-driven light source (LDLS) was coupled into the optical fibers, and passed through the collimators and a 450 nm-centered bandpass filter (Thorlabs FB 450-40, full-width at half-maximum (FWHM) of 40 nm). Thus, light of 410–490 nm was entered into the optical cavity. The cavity includes integrating sphere, truncation reduction tubes and highly reflective (HR) mirrors. The sphere manufactured with aluminum was separated into two hemispheres, and each hemisphere has a hole for the probe light beam. Moreover, a hole in the side wall is used to measure the scattered signal. SOA particles were introduced into the optical cavity via the forward truncation reduction tube, and then by encountering and interacting with the light in integrating sphere, the generated scattering signal was detected with a photomultiplier tube (PMT). A light baffle was utilized to eliminate the stray light, and a 470 nm-centered (FWHM of 9 nm) bandpass filter was used to prevent the scattered light from directly reaching the PMT. The PMT signal was acquired with a data acquisition card, which provided an integrated scattering signal over 465–474 nm. Meanwhile, the exited light was coupled into the fiber and sent to the CCD spectrometer to measure the extinction signal. The scattering and extinction coefficient at 470 nm was deduced as an averaged scattering and extinction value in the range of 465–474 nm, and utilized for data analysis [14].

2.4. Characterization Compositions of Toluene SOA

After 4 h photooxidation, the toluene SOA particles were collected onto the polytetra fluoroethylene membrane filter and extracted into 5 mL 2% methanol water solution with 30 min sonication [25]. A secondary extraction was performed to confirm that the primary extraction was complete when the absorption spectrum of its extraction is approximately a straight line. The extraction was filtrated with syringe organic filter (33 mm × 0.22 um, Millipore Corp, Billerica, MA, USA) before the UV-Vis and liquid chromatography-mass spectrometer (LC-MS) measurements. The double beam UV-6100S spectrophotometer (Mapada Instruments, Shanghai, China) was utilized to determine the ultraviolet-visible spectrum of extract solution, with 2% methanol water solution as the reference solution. Moreover, extract solutions were measured by LC-MS with electrospray ionization (ESI) (Agilent Technologies Inc, Palo Alto, CA, USA, Model 1200 and 6320). The autosampler injected the extract sample (20 μL) into the LC system, which introduced the sample into the ESI source region. The extract sample was detected without LC column. The mobile phase was ultra-pure water and methanol (1:1 v/v), with the rate of 0.20 mL/min. The extract solutions were analyzed in 50–1000 amu of the negative mode with a fragmentor voltage of 40 V and a capillary voltage of 3000 V. N2 was the drying gas (350 °C, 24 psig, 10 L/min). The mass spectra were recorded on Agilent software (Chemstation Rev.b.01.03) and exported to Excel for statistical analysis and interpretation [25].
It should be noted that we have not phased out the effect of salt in UV-Vis and subsequent MS measurements of the toluene-SOA extraction. According to the experimental results of Bone et al. [26], imidazoles formed via the aqueous reaction with limonene SOA and NH4+ ion over days. Toluene SOA particles were extracted and detected by UV-Vis and LC-MS immediately, so the effect of (NH4)2SO4 on the aqueous reaction was negligible. In addition, the m/z of sulfate was 48, and the toluene-SOA extraction was detected in 50–1000 amu of the negative mode, thus, (NH4)2SO4 in the solution would not affect the measurement of chemical components of SOA.

3. Results

3.1. Validation of the Retrieved CRI from the Albedometer

The accuracy and precision of the retrieved CRI, scattering and extinction coefficient from the cavity-enhanced albedometer were tested and evaluated using laboratory-generated mono-disperse polystyrene latex (PSL) particles and polydisperse ambient particles in our previous published papers [13,14,15]. PSL particles were generated with the atomizer (TSI Inc, Shoreview, MN, USA, Model 3076), and 200, 240, 300 and 400 nm particles were selected by the classifier (TSI Inc, Shoreview, MN, USA, Model 3080L) and supplied to the albedometer to measure the scattering (Iscat) and extinction (Itrans) signal, respectively. It should be noted that a small number of multiplied charged particles emerged from the differential mobility analyzer (DMA) selector. These particles were characterized by the tandem DMA method suggested by Bueno et al. [27], and the error of aerosol extinction cross section caused by multiply charged particles was estimated to be 5–20% for the particle diameters smaller than 300 nm, and less than 5% for the particle diameters larger than 400 nm [13]. As suggested by Washenfelder et al. [28] and Thompson et al. [29], the extinction (αext) and scattering (αscat) coefficient can be calculated from the measured Iscat and Itrans. The CRI can be retrieved from the simultaneous measurement of αext and αscat coefficients, subsequently, and the detailed retrieval algorithm was presented in our previous works [13,14]. The retrieved CRI of PSL particles was m = 1.676 + i 0.015 via the scattering signal and m = 1.674 + i 0 via the extinction signal at λ = 470 nm, which agreed with the reported result of 1.633 + i 0.005 at λ = 420 nm from Washenfelder et al. [28], and 1.627 + i 0.005 at λ = 560 nm provided by Miles et al. [30] However, the imaginary part value of PSL particles retrieved from the scattering channel was larger than these reported values, which may be due to the large internal volume, the longer residual time and the greater agglomeration effect on small particles. The potential uncertainty for the retrieved CRI value was about 5% and 3% via the scattering and extinction signal, respectively [14].
Polydisperse ambient particles measurements were performed outside the laboratory for 24 h. The scattering, extinction coefficients and SSA of the ambient air were measured with the aerosol albedometer, and the scattering coefficients are compared with the measurements from the TSI 3563 integrating nephelometer. A good agreement between the albedometer and the TSI nephelometer is observed [14], demonstrating that albedometer can measure αext, αscat and SSA of polydisperse aerosol particles simultaneously. Simultaneous measurement of αext and αscat coefficients provides an approach for retrieval of CRI. The detailed retrieval algorithm has been presented in our previous paper [15]. The total uncertainty for the retrieved real CRI value of polydisperse aerosol particles was estimated to be less than 3% and 2% via the scattering and extinction signal, respectively. Additionally, the relative uncertainty for the imaginary part of CRI was estimated to be between 15% and 20% for the k-value less than 0.010, and less than 7% for the k-value greater than 0.010 from the scattering channel, while the uncertainty of the associated imaginary CRI value was better than 10 % for the k-value less than 0.010, and less than 5% for the k-value greater than 0.010 through the extinction channel. The total uncertainty in the measurement of SSA was estimated to be less than 5% [13,14,15].

3.2. Optics of SOA without and in Presence of (NH4)2SO4 Seed

The mass concentration and optics of SOA without (NH4)2SO4 seed were measured firstly. Unless mentioned otherwise, the optical properties of SOA were derived at 470 nm. Similar to our previous work, the density of 1.23 g/cm3 was used to estimate the mass concentration of particle measured by SMPS [31]. Wall loss was described as a first order process dependent on the loss coefficient, kdep (dp):
k d e p ( d p ) = a d p b + c / d p d
where a, b, c, and d were optimized to be 4.17 × 10−13, 4.66, 10.18, and 0.75, respectively [22,31]. The mass concentration of SOA was corrected for wall loss by fitting the particle number concentration decay at the end of the experiment. For the mixed gases of 600 ppb toluene and 3000 ppb H2O2, the corrected mass concentration, particle number concentration, mean diameter, SSA and other optical parameters of toluene SOA with different reaction time measured by SMPS and albedometer are shown in Figure 2 and Figure 3. In the first 5 min of the photooxidation time, only few particles with a corrected mass concentration of 1.5 μg/m3 was observed. OH radicals generated from the photolysis of H2O2 undergo hydrogen extraction and addition reactions with toluene to form semi- and non-volatile products. According to the theory of gas/particle partitioning [32], these gaseous products nucleate homogeneously only after exceeding their saturation concentration. So, nearly no SOA particles would be formed until sufficient toluene had reacted to generate gaseous concentrations that exceed saturation concentration. Meanwhile, afterward, the fine particles less than 100 nm increased sharply. Within 5–35 min, the particle number concentration of toluene SOA increased from ~200 pt/cm3 to ~150,000 pt/cm3, the corrected mass concentration of SOA increased from 1.5 μg/m3 to 74 μg/m3, and the scattering and extinction coefficient measured by the albedometer increased to about 450 and 480 Mm−1, respectively. In the following 35–65 min, the particle number concentration of SOA decreases and the particle size increases gradually, due to more gaseous products condensed on the pre-exiting particles and collision between fine particles. At about 65 min, the particle number concentration of toluene SOA decreases to ~100,000 pt/cm3, the mean diameter increases to 240 nm, the corrected mass concentration of SOA increases to 145 μg/m3, and αscat and αext reach the maximum value of 680 and 740 M m−1, respectively. Thereafter, no toluene react to form new particles; the size distribution of SOA remained basically unchanged. However, due to turbulence and gravity deposition, toluene SOA particles deposited on the chamber wall [22,23], resulting in a gradual decrease in the number concentration, scattering coefficient and absorption coefficient of toluene SOA particles, as illustrated in Figure 3. As the wall loss is the main factor in reducing the mass concentration of SOA, the corrected mass concentration of SOA was almost constant after 65 min, as shown in Figure 2.
It can be seen from Figure 3 that αext, αscat and SSA of SOA increase with the increasing of particle size. According to the measurement results of GC, toluene in the chamber has been almost completely consumed after 65 min of illumination, after that the mean diameter and size distribution of SOA were maintained almost constantly. Although the wall effect causes the particle number concentration, extinction and scattering coefficient of toluene SOA to decrease gradually, SSA remains substantially unchanged. SSA characterizes the relative strength of the scattering and absorption. Its value ranges from 1 for purely scattering to 0 for completely absorbing particle [6]. The SSA of SOA was measured to be 0.92 ± 0.02, indicating that toluene SOA particles without (NH4)2SO4 seed have strong scattering ability. This was further confirmed by the retrieved CRI from the scattering and extinction channel. The retrieved real and imaginary part of the CRI of the toluene SOA is 1.486 ± 0.002 and 0.006 ± 0.001 via the scattering signal, and 1.479 ± 0.002 and 0.004 ± 0.001 via the extinction signal, respectively, when the size distribution of SOA is kept nearly unchanged. The real CRIs of toluene SOA retrieved from scattering and extinction channels fall well within the range of 1.431–1.498 at λ = 532 nm reported by Nakayama et al. [8] and 1.450–1.518 at λ = 532 nm measured by Li et al. [9]. It is worth noting that the CRI retrieved by the scattering channel is slightly larger than that obtained by the extinction channel. Since the real CRI of toluene obtained by the scattering channel is closer to the corresponding value (1.518) of toluene SOA with low-NOx and H2O2 was utilized as an OH precursor performed by Li et al. [9], the CRI retrieved by the scattering channel was shown in Figure 3 and used to investigate the effect of (NH4)2SO4 seed on the optical properties of toluene SOA. The imaginary part k is the characterization of the ability of absorbing light radiation. The larger the k value, the stronger the light absorption of aerosol particles. The measured k-value of toluene SOA particles shown in Figure 3 is only 0.006 ± 0.001, indicating that toluene SOA without (NH4)2SO4 seed has weak absorb ability and contributes mainly to a cooling effect.
The SSA and CRI of SOA particles formed in the presence of different concentration of (NH4)2SO4 seed were measured to explore the influences of (NH4)2SO4 seed on the optics of toluene SOA. For the (NH4)2SO4 seeded experiments, the corrected mass concentration of SOA was calculated by subtracting the initial seed mass concentration from the wall loss corrected aerosol mass concentration. The obtained curves of the corrected mass concentration, particle number concentration, mean diameter, extinction and scattering coefficients, SSA and CRI of toluene SOA particles formed in the presence of 300 μg/m3 (NH4)2SO4 seed with different reaction time displayed in Figure 2 and Figure 4 are similar to the situation without (NH4)2SO4 seed. However, the corrected maximum concentration of toluene SOA reached 242 μg/m3 at 65 min, which was 67% higher than that without seed aerosol, indicating that (NH4)2SO4 seed aerosol can promote the formation of toluene SOA. Different from the case without (NH4)2SO4 seed where the extinction coefficient is approximately equal to the scattering coefficient, the measured extinction coefficient of toluene SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed is larger than that of scattering coefficient after 65 min of illumination. The measured SSA illustrated in Figure 4 was 0.81 ± 0.02, less than the SSA of toluene SOA particles without (NH4)2SO4 seed of 0.92, and slightly lower than SSA of biomass burning organic aerosols range in 0.84–0.93 [33], indicating that the toluene SOA formed in the presence of (NH4)2SO4 seed had a certain light-absorption capacity.
As shown in Figure 3 and Figure 4, the real CRI of toluene SOA formed in the absence and presence of 300 μg/m3 (NH4)2SO4 seed decreased as time went on. The real part value of CRI for toluene SOA without seed was found to decrease from 1.608 to 1.486 when CRI was nearly stable. Furthermore, the real CRI of SOA formed in presence of (NH4)2SO4 seed was smaller than that of without seed, and its real CRI decrease from 1.592 to 1.285. Similar to the system of m-xylene SOA performed by Li et al. [10], the relatively low volatile gaseous product of photooxidation of toluene nucleated to generate SOA particles in the absence of seed. As the reaction proceeded, the concentration of organics increased, particles got larger, and volatile products with low CRI would condense on the particles, which led to the real CRI of SOA to decrease with time, as shown in Figure 3. The presence of (NH4)2SO4 seed promoted the condensation of volatile organics [10,34], condensation of gaseous products of photooxidation of toluene and further heterogeneous reactions on the surface of seed can possibly form a core-shell mixing state at RH = 25%. More and more higher volatility products with lower CRI condensed on the particles as time gone on, and the real CRI of SOA formed in presence of 300 μg/m3 (NH4)2SO4 seed displayed in Figure 4 decreased significantly when compared to that without seed.
Figure 5 displayed the corrected maximum mass concentration of tolunene SOA detected by SMPS, SSA, the real (n) and imaginary (k) part of CRI of SOA measured at different concentrations of (NH4)2SO4 seed when the size distribution of SOA particles remains basically unchanged. As shown in Figure 5a, the corrected maximum concentration of tolunene SOA at different concentrations of (NH4)2SO4 seed is in the range of 166–242 μg/m3, which is higher than that without seed (145 μg/m3). It should be noted that, when the concentration of (NH4)2SO4 seed aerosol was greater than 200 μg/m3, the maximum concentration of toluene SOA did not continue to increase. SSA and real CRI of SOA decrease, while k of SOA increases with the increasing concentration of (NH4)2SO4 seed (Figure 5b–d). Similarly, when (NH4)2SO4 seed exceeds 200 μg/m3, the optical parameters of SOA tend to be stable. The n-value of toluene SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed is 1.285, 14% less than that of without seed (1.486), and the k-value of SOA is 0.042, 6 times larger than SOA generated in absence of (NH4)2SO4 seed of 0.006, demonstrating that (NH4)2SO4 seed is involved in the photochemical reaction of toluene with OH radicals, resulting in products with light absorbing ability.

3.3. Components of Toluene SOA without and in Presence of (NH4)2SO4 Seed

The OH-initiated photooxidation of toluene generates methyl-hydroxy-cyclohexadienyl and benzyl radical radical via OH addition and hydrogen abstraction, respectively [35,36]. As shown in Figure 6, the reactions of the benzyl radical with O2 led to the generation of benzaldehyde, and the methyl-hydroxy-cyclohexadienyl radical reacts with O2 by H-abstraction to form cresol and O2 addition to yield peroxy radical, which undergo a series of reactions to yield 5-methyl-6-oxo-2,4- hexadienal, glyoxal, and other aldehydes. Furthermore, aldehyde products, such as glyoxal, methylglyoxal can further be oxidized to carboxylic acids [37]. As proposed by Suh et al. [38], the reaction channel of methylhydroxycyclohexadienyl radical with O2 by hydrogen abstraction had been shown to be relatively minor, demonstrating that carboxyl compounds are the principal components of toluene SOA in the absence of (NH4)2SO4 seed. These results are further confirmed by the ESI-MS and UV-Vis spectra of SOA without (NH4)2SO4 seed illustrated in Figure 7 and Figure 8.
Due to the small volume of our chamber (850 L), coupled with wall effects and other effects, the mass of the collected particulate matter is only tens to hundreds of micrograms, and we cannot accurately measure its mass with the electronic balance (accurate to 0.0001 g). The UV-Vis absorption spectrum was used to qualitatively verify the component of toluene SOA. Therefore, we estimated the mass of the SOA collected on the filter, based on the concentration of the particles measured by SMPS at the end of the reaction. The concentration of toluene SOA without seed at the end of the reaction is about 60 μg/m3, the remaining volume in the chamber is about 700 L, and the mass of toluene SOA collected is estimated to be about 40 μg; the concentration of extract for SOA is about 8 mg/L. As suggested by Carlton et al. [39], ESI deprotonated compounds and the formed negative ions are molecular-related ions [M–H], which provided the information of organic’s molecular weight. The [M–H] ions of carboxyl compounds, such as glyoxal (m/z 57), methylglyoxal (m/z 71), methylglyoxylic acid (m/z 87), oxalic acid (m/z 89), 4-oxo-2-pentenoic acid (m/z 113) and benzoic acid (m/z 121), 2-methyl-4-oxo-2,3-epoxy-butyric acid (m/z 129), 2-methyl-2,3-epoxy-succinic acid (m/z 139), 5- methyl-6-oxo-2,4-hexadienoic acid (m/z 145) and 5-methyl-2,4-hexadiendioic acid (m/z 155) appear in the ESI-MS shown in Figure 7. In addition, the absorption band at 205 nm of characteristic light absorption of carboxyl compounds [38] emerged in the ultraviolet-visible spectrum of SOA, as displayed in Figure 8a. The formed aldehydes and carboxylic acids contain only C=C and C=O double bonds, and do not contain strong chromophores and auxochromes [40]. Thus, toluene SOA particles without (NH4)2SO4 seed have weak light absorption.
Toluene SOA concentration formed in the presence of 300 μg/m3 (NH4)2SO4 seed at the end of the reaction is about 340 μg/m3, and the mass of collected particles is estimated to be about 240 μg. In order to estimate the mass of the collected toluene SOA formed in the presence of seed, 300 μg/m3 (NH4)2SO4 seed was separately prepared for photooxition reaction. After 4 hours’ photooxidation, the particle concentration was about 250 μg/m3, and the mass of collected ammonium sulfate was estimated to be about 175 μg. Thus, the mass of collected toluene SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed was about 65 μg, and the concentration of extract for SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed was about 13 mg/L. As suggested by Updyke et al. [41], the mass absorption coefficient (MAC, m2/g) can be calculated from the absorbance A(λ) of the SOA extract, with solution mass concentration Cmass (g/m3) measured over pathlength b (m):
M A C ( λ ) = A ( λ ) × ln 10 b × C m a s s
Figure 8b gave absorbance of 0.0045 at 470 nm, and the corresponding MAC was estimated to be 0.080 m2/g, which was comparable to MAC values of brown carbon produced by aging SOA with NH3 performed by Updyke et al. [41] This indicated that the measured UV-Vis spectrum can reflect the optical characteristics of toluene SOA.
The UV-Vis absorption profile of toluene SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed shown in Figure 8b is different from that of without seed. There is a distinctive band at 280 nm in spectra of toluene SOA formed in presence of (NH4)2SO4 seed, indicating the generation of new products. The band of 280 nm was also emerged in the spectra of aqueous reaction products of (NH4)2SO4 and glyoxal measured by Kampf et al. [42], Lee et al. [43] and Maxut et al. [44] They considered that the chromophores were probably imidazole products formed from the reactions between NH4+ and glyoxal, and the n→π* transition in the C=N of imidazoles was the main contributor to the 280 nm band. As glyoxal and methylglyoxal are the principal α-dicarbonyl gaseous products formed by photooxidation of toluene [36,45], similar reactions to form imidazole compounds may occur after α-dicarbonyls condensate on the surface of (NH4)2SO4 seed. Furthermore, the obtained negative mass spectra of toluene SOA formed in the presence of 300 μg /m3 (NH4)2SO4 seed illustrated in Figure 9 contained [M–H] of CxHyNnOz family of imidazole derivative compounds. In addition, [M–H] of 67, 81, 95, 123 could be attributed to C3H3N2, C4H5N2, C4H3N2O and C6H4N2O, respectively, which were also detected by Liu et al. [46] in the experiment of heterogeneous uptake of NH3 by m-xylene SOA and the previous study of aged benzene SOA formed in presence of (NH4)2SO4 seed [23]. These results indicated that imidazoles are the newly formed particulate products of toluene SOA in the presence of (NH4)2SO4 seed.
According to the experimental results of Trainic et al. [47], heterogeneous reactions between glyoxal and dehydrated ammonium sulfate cannot occur at RH of 35%. Compared with the reaction system of glyoxal and ammonium sulfate seed performed by Trainic et al. [47], the composition of our reaction system is more complex. In addition to glyoxal, carboxylic acids such as formic acid, acetic acid and glyoxylic acid were also produced from the photooxidation of toluene [36,45]. Some studies have reported that the presence of organics in SOA-coated (NH4)2SO4 particles could increase the particles’ water uptake compared to that of pure (NH4)2SO4 particles, even with RH less than 30% [48,49,50]. Meyer et al. [50] found that SOA coating enhanced the water uptake of (NH4)2SO4 seed, and led to a partially dissolved (NH4)2SO4 seed. Thus, (NH4)2SO4 seed can partially be hydrated by absorbing water from its covered outer organic layer, with RH of 25% of our reaction system. Additionally, the gaseous carboxylic acid products of toluene condense on the water layer of the seed to produce acidic aqueous layer for heterogeneous imidazole formation [22]. The gas/particle partitioning of degradation products on seed particle are principal contributions to the formation of toluene SOA formed in presence of (NH4)2SO4 seed [46]. When glyoxal, methylglyoxal and other α- dicarbonyls are partitioned on the (NH4)2SO4 seed, heterogeneous acid-catalyzed reactions occur to form imidazole products. Glyoxal and methylglyoxal can be protonated by H+ ion and hydrolyzed to tetrol product (1), as displayed in Figure 9. Additionally, the protonated α-dicarbonyls can react with NH4+ ions to produce diimine product (2). Tetrol product (1) can react with diimine (2) to form (3) through the dehydration reaction. However, (3) is unstable, N atom attacks C atom, generating (4) after dehydration. Moreover, (4) forms HCOOH (or CH3COOH) and imidazole (m/z 67) (or 4-methyl-imidazole (m/z 81)) via rearrangement and dehydration reaction. Additionally, (4) occurs rearrangement illustrated in Figure 10 to produce (5), which could be dehydrated to yield imidazole-2-carbaldehyde (m/z 95) (or 4-methyl-imidazole-2- acetaldehyde (m/z 123)). As depicted in Figure 10, the formed imidazole (or 4-methyl-imidazole) can subsequently interact with (1) to produce hydrated N-glyoxal substituted imidazole (m/z 143) (or hydrated N-methylglyoxal substituted 4-methyl-imidazole (m/z 171)), hydrated glyoxal dimer substituted imidazole (m/z 201) (or hydrated methylglyoxal dimer substituted 4-methyl-imidazole (m/z 243)), respectively [42,43,44].
The high concentration of (NH4)2SO4 seed provides more surface area for condensation and reaction of gaseous α-dicarbonyls. Thus, the content of imidazoles in SOA increases with the concentration of seed aerosol, which leads to the increase of absorptivity of SOA with the concentration of seed aerosol, as shown in Figure 5. It is noteworthy that the concentration of toluene, H2O2 and other conditions are basically unchanged in all experiments, and the amount of the formed α-dicarbonyls is also nearly constant for each experiment. When (NH4)2SO4 seed is increased to a certain concentration (200 μg/m3), all the α-dicarbonyls are consumed completely, and the formation of imidazoles would not increase with the increasing concentration of (NH4)2SO4 seed. Thus, the optical properties of toluene SOA remain basically unchanged when (NH4)2SO4 seed exceeds 200 μg/m3, as illustrated in Figure 5.
It should be pointed out that, due to the low sensitivity of UV-Vis spectrophotometer with 1 cm optical length, almost no absorption of toluene SOA is measured at 380 nm, as shown in Figure 8. However, the albedometer with 3.7 km effective optical path can detect the absorption of toluene SOA with 465–474 nm [14]. Compared to the studies of Nakayama et al. [7,8] and Li et al. [9,10], the present study extended (NH4)2SO4 seed to 300 μg/m3, and the optics of SOA were detected by aerosol single-scattering albedometer in real-time. The measured SSA of toluene SOA formed in the presence of 300 μg/m3 (NH4)2SO4 seed was 0.81 ± 0.02, showing a certain light-absorption capacity [6,40]. Additionally, the SSA of toluene SOA was observed to decrease from 0.91 to 0.81, while the imaginary part of CRI (k) was found to increase from 0.013 to 0.042 with increasing the (NH4)2SO4 seed concentration from 25 to 300 μg/m3, demonstrating that the adsorption capacity of toluene derived SOA increased with the increasing concentration of (NH4)2SO4 seed. From absorption and mass spectrum of SOA, we confirmed that the decrease of SSA and increase of k caused by (NH4)2SO4 seed are due to the increase of imidazoles in SOA. The C=N chromophores of imidazoles enhance the absorption capacity of SOA. According to the UV-Vis absorption spectra shown in Figure 8, toluene SOA formed in the presence of (NH4)2SO4 seed has strong absorption ability in the UV range. Owing to the rate of global production of anthropogenic SOA is larger than biomass burning organic aerosols [51]; aromatic SOA in the urban atmosphere containing high concentrations of (NH4)2SO4 fine particles may contribute remarkably to climate forcing.

4. Conclusions

The aromatic SOA in the urban atmosphere with high concentrations of (NH4)2SO4 seed are considered as a potential major component of brown carbon. The toluene SOA particles were formed with different concentrations of (NH4)2SO4 seed in the chamber, and measured by the aerosol single-scattering albedometer, spectroscopic and mass spectrometer in the present study. The detected SSA of SOA formed in the presence of (NH4)2SO4 seed is less than the SOA without seed, and the imaginary part of CRI (k) was found to increase with increasing the concentration of (NH4)2SO4 seed. Compared with the SOA without seed, imidazoles inferred from absorption and mass spectra of SOA formed in the presence of (NH4)2SO4 seed were confirmed to be responsible for the decreased SSA and increased k. As the SSA of SOA formed in presence of (NH4)2SO4 seed is lower than BBOA detected at λ = 550 nm, while absorption capacity in the ultraviolet range is higher and the rate of production of SOA is larger than BBOA, the anthropogenic SOA with high concentration of (NH4)2SO4 fine particles may contribute to the radiative balance in urban atmosphere. Although the concentration of (NH4)2SO4 seed in the chamber is higher than actual atmosphere, the influence of high mass concentration of (NH4)2SO4 seed is still expected to be prevalent in Chinese Urban atmosphere as organic aerosol mass loading is low, under which condition the incidence of mass concentration of (NH4)2SO4 effect could be high [52]. These experimental results could be utilized to improve the estimation of radiative forcing of anthropogenic SOA in some regions. Nevertheless, the components of SOA were qualitatively measured by UV-Vis and LC-MS; a suitable chromatographic column should be selected to separate and quantitatively analyze the constituents of toluene SOA. Additionally, SOA were generated in the presence of (NH4)2SO4 seed with a relative humidity (RH) of 25 ± 2%; the effects of RH on the optics of aromatic SOA should be carried out in future experiments.

Author Contributions

Conceptualization, M.H. and W.Z. (Weijun Zhang); methodology, W.Z. (Weijun Zhang) and W.Z. (Weixiong Zhao); software, X.G.; validation, M.H. and W.Z. (Weijun Zhang); formal analysis, W.Z. (Weixiong Zhao) and C.H.; data curation, T.L.; writing—original draft preparation, T.L.; writing—review and editing, M.H.; project administration, M.H.; funding acquisition, M.H. and C.H. All authors reviewed and commented on the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 41575118, 41575126, 41305109), the Key Project of Foundation of Fujian Province of China (No.2020J06018), the science and technology project of Minnan Normal University (No. 4201-L11805).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Drozd, G.T.; Zhao, Y.; Saliba, G.; Frodin, B.; Maddox, C.; Chang, M.-C.O.; Maldonado, H.; Sarder, S.; Weber, R.J.; Robinson, A.L.; et al. Detailed Speciation of Intermediate Volatility and Semivolatile Organic Compound Emissions from Gasoline Vehicles: Effects of Cold-Starts and Implications for Secondary Organic Aerosol Formation. Environ. Sci. Technol. 2018, 53, 1706–1714. [Google Scholar] [CrossRef] [PubMed]
  2. Derwent, R.G. Representing Organic Compound Oxidation in Chemical Mechanisms for Policy-Relevant Air Quality Models under Background Troposphere Conditions. Atmosphere 2020, 11, 171. [Google Scholar] [CrossRef] [Green Version]
  3. Baltensperger, U. Aerosols in Clearer Focus. Science 2010, 329, 1474–1475. [Google Scholar] [CrossRef] [PubMed]
  4. Donateo, A.; Feudo, T.L.; Marinoni, A.; Dinoi, A.; Avolio, E.; Merico, E.; Calidonna, C.R.; Contini, D.; Bonasoni, P. Characterization of In Situ Aerosol Optical Properties at Three Observatories in the Central Mediterranean. Atmosphere 2018, 9, 369. [Google Scholar] [CrossRef] [Green Version]
  5. Chen, N.; Zhao, Y.; Lyu, R.; Wu, R.; Dai, L.; Zhao, Y.; Chen, F.; Zhang, J.; Yu, H.; Guan, M. Seasonal and spatial variations of optical properties of light absorbing carbon and its influencing factors in a typical polluted city in Yangtze River Delta, China. Atmos. Environ. 2019, 199, 45–54. [Google Scholar] [CrossRef]
  6. Moise, T.; Flores, J.M.; Rudich, Y. Optical Properties of Secondary Organic Aerosols and Their Changes by Chemical Processes. Chem. Rev. 2015, 115, 4400–4439. [Google Scholar] [CrossRef]
  7. Nakayama, T.; Matsumi, Y.; Sato, K.; Imamura, T.; Yamazaki, A.; Uchiyama, A. Laboratory studies on optical properties of secondary organic aerosols generated during the photooxidation of toluene and the ozonolysis of α-pinene. J. Geophys. Res. Space Phys. 2010, 115. [Google Scholar] [CrossRef]
  8. Nakayama, T.; Sato, K.; Matsumi, Y.; Imamura, T.; Yamazaki, A.; Uchiyama, A. Wavelength and NOx dependent complex refractive index of SOAs generated from the photooxidation of toluene. Atmos. Chem. Phys. 2013, 13, 531–545. [Google Scholar] [CrossRef] [Green Version]
  9. Li, K.; Wang, W.; Ge, M.F.; Li, J.; Wang, D. Optical properties of secondary organic aerosols generated by photooxidation of aromatic hydrocarbons. Sci. Rep. 2014, 4, 4922. [Google Scholar] [CrossRef]
  10. Li, K.; Li, J.; Wang, W.; Li, J.; Peng, C.; Wang, D.; Ge, M. Effects of Gas-Particle Partitioning on Refractive Index and Chemical Composition of m-Xylene Secondary Organic Aerosol. J. Phys. Chem. A 2018, 122, 3250–3260. [Google Scholar] [CrossRef]
  11. Langridge, J.M.; Richardson, M.S.; Lack, D.A.; Brock, C.A.; Murphy, D.M. Limitations of the Photoacoustic Technique for Aerosol Absorption Measurement at High Relative Humidity. Aerosol Sci. Technol. 2013, 47, 1163–1173. [Google Scholar] [CrossRef] [Green Version]
  12. Varma, R.M.; Ball, S.M.; Brauers, T.; Dorn, H.-P.; Heitmann, U.; Jones, R.L.; Platt, U.; Pöhler, D.; Ruth, A.A.; Shillings, A.J.L.; et al. Light extinction by secondary organic aerosol: An intercomparison of three broadband cavity spectrometers. Atmos. Meas. Tech. 2013, 6, 6685–6727. [Google Scholar] [CrossRef]
  13. Zhao, W.X.; Dong, M.L.; Chen, W.D.; Gu, X.J.; Hu, C.J.; Gao, X.M.; Huang, W.; Zhang, W.J. Wavelength resolved optical extinction measurements of aerosols using broad-band cavity-enhanced absorption spectroscopy over the spectral range of 445–480 nm. Anal. Chem. 2013, 85, 2260–2268. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, W.; Xu, X.; Dong, M.; Chen, W.; Gu, X.; Hu, C.; Huang, Y.; Gao, X.; Huang, W.; Zhang, W. Development of a cavity-enhanced aerosol albedometer. Atmos. Meas. Tech. 2014, 7, 2551–2566. [Google Scholar] [CrossRef] [Green Version]
  15. Xu, X.; Zhao, W.; Zhang, Q.; Wang, S.; Fang, B.; Chen, W.; Venables, D.S.; Wang, X.; Pu, W.; Wang, X.; et al. Optical properties of atmospheric fine particles near Beijing during the HOPE-J3A campaign. Atmos. Chem. Phys. 2016, 16, 6421–6439. [Google Scholar] [CrossRef] [Green Version]
  16. Wang, S.; Zhao, W.; Xu, X.; Fang, B.; Zhang, Q.; Qian, X.; Zhang, W.; Chen, W.; Pu, W.; Wang, X. Dependence of columnar aerosol size distribution, optical properties, and chemical components on regional transport in Beijing. Atmos. Environ. 2017, 169, 128–139. [Google Scholar] [CrossRef]
  17. Ge, S.; Xu, Y.; Jia, L. Effects of inorganic seeds on secondary organic aerosol formation from photochemical oxidation of acetone in a chamber. Atmos. Environ. 2017, 170, 205–215. [Google Scholar] [CrossRef]
  18. Tajuelo, M.; Rodríguez, A.M.; Baeza-Romero, M.T.T.; Aranda, A.; Díaz-De-Mera, Y.; Rodríguez, D. Secondary organic aerosol formation from α-methylstyrene atmospheric degradation: Role of NO level, relative humidity and inorganic seed aerosol. Atmos. Res. 2019, 230, 104631. [Google Scholar] [CrossRef]
  19. Shao, P.; Tian, H.; Sun, Y.; Liu, H.; Wu, B.; Liu, S.; Liu, X.; Wu, Y.; Liang, W.; Wang, Y.; et al. Characterizing remarkable changes of severe haze events and chemical compositions in multi-size airborne particles (PM1, PM2.5 and PM10) from January 2013 to 2016–2017 winter in Beijing, China. Atmos. Environ. 2018, 189, 133–144. [Google Scholar] [CrossRef]
  20. Wang, J.; Zhang, J.; Liu, Z.-J.; Wu, J.; Zhang, Y.-F.; Han, S.-Q.; Zheng, X.-J.; Zhou, L.-D.; Feng, Y.-C.; Zhu, T. Characterization of chemical compositions in size-segregated atmospheric particles during severe haze episodes in three mega-cities of China. Atmos. Res. 2017, 187, 138–146. [Google Scholar] [CrossRef]
  21. Sun, Z.; Duan, F.; He, K.; Du, J.; Zhu, L. Sulfate–nitrate–ammonium as double salts in PM2.5: Direct observations and implications for haze events. Sci. Total. Environ. 2019, 647, 204–209. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, M.; Hao, L.; Cai, S.; Gu, X.; Zhang, W.; Hu, C.; Wang, Z.; Fang, L.; Zhang, W. Effects of inorganic seed aerosols on the particulate products of aged 1,3,5-trimethylbenzene secondary organic aerosol. Atmos. Environ. 2017, 152, 490–502. [Google Scholar] [CrossRef]
  23. Huang, M.; Zhang, W.; Cai, S.; Liao, Y.; Zhao, W.; Hu, C.; Gu, X.; Fang, L.; Zhang, W. Mass spectrometric study of aged benzene secondary organic aerosol in the presence of dry ammonium sulfate. J. Atmos. Chem. 2016, 73, 329–344. [Google Scholar] [CrossRef]
  24. Assaf, E.; Fittschen, C. Cross section of OH radical overtone transition near 7028 cm–1 and measurement of the rate constant of the reaction of OH with HO2 radicals. J. Phys. Chem. A 2016, 120, 7051–7059. [Google Scholar] [CrossRef] [PubMed]
  25. Feng, Z.; Huang, M.; Cai, S.; Xu, X.; Yang, Z.; Zhao, W.; Hu, C.; Gu, X.; Zhang, W. Characterization of single scattering albedo and chemical components of aged toluene secondary organic aerosol. Atmos. Pollut. Res. 2019, 10, 1736–1744. [Google Scholar] [CrossRef]
  26. Bones, D.L.; Henricksen, D.K.; Mang, S.A.; Gonsior, M.; Bateman, A.P.; Nguyen, T.B.; Cooper, W.J.; Nizkorodov, S.A. Appearance of strong absorbers and fluorophores in limonene-O3 secondary organic aerosol due to NH4+-mediated chemical aging over long time scales. J. Geophys. Res. Space Phys. 2010, 115. [Google Scholar] [CrossRef] [Green Version]
  27. Bueno, P.A.; Havey, D.K.; Mulholland, G.W.; Hodges, J.T.; Gillis, K.A.; Dickerson, R.R.; Zachariah, M.R. Photoacoustic Measurements of Amplification of the Absorption Cross Section for Coated Soot Aerosols. Aerosol Sci. Technol. 2011, 45, 1217–1230. [Google Scholar] [CrossRef]
  28. Washenfelder, R.A.; Flores, J.M.; Brock, C.A.; Brown, S.S.; Rudich, Y. Broadband measurements of aerosol extinction in the ultraviolet spectral region. Atmos. Meas. Tech. 2013, 6, 861–877. [Google Scholar] [CrossRef] [Green Version]
  29. Thompson, J.E.; Barta, N.; Policarpio, D.; Duvall, R. A fixed frequency aerosol albedometer. Opt. Express 2008, 16, 2191–2205. [Google Scholar] [CrossRef]
  30. Miles, R.E.H.; Rudić, S.; Orr-Ewing, A.J.; Reid, J.P. Influence of Uncertainties in the Diameter and Refractive Index of Calibration Polystyrene Beads on the Retrieval of Aerosol Optical Properties Using Cavity Ring Down Spectroscopy. J. Phys. Chem. A 2010, 114, 7077–7084. [Google Scholar] [CrossRef]
  31. Huang, M.; Hao, L.; Gu, X.; Hu, C.; Zhao, W.; Wang, Z.; Fang, L.; Zhang, W. Effects of inorganic seed aerosols on the growth and chemical composition of secondary organic aerosol formed from OH-initiated oxidation of toluene. J. Atmos. Chem. 2013, 70, 151–164. [Google Scholar] [CrossRef]
  32. Jathar, S.H.; Mahmud, A.; Barsanti, K.C.; Asher, W.E.; Pankow, J.F.; Kleeman, M.J. Water uptake by organic aerosol and its influence on gas/particle partitioning of secondary organic aerosol in the United States. Atmos. Environ. 2016, 129, 142–154. [Google Scholar] [CrossRef] [Green Version]
  33. Reid, J.S.; Eck, T.F.; Christopher, S.A.; Koppmann, R.; Dubovik, O.; Eleuterio, D.P.; Holben, B.N.; Reid, E.A.; Zhang, J. A review of biomass burning emissions part III: Intensive optical properties of biomass burning particles. Atmos. Chem. Phys. 2005, 5, 827–849. [Google Scholar] [CrossRef] [Green Version]
  34. 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.; et al. Effect of oxidant concentration, exposure time, and seed particles on secondary organic aerosol chemical composition and yield. Atmos. Chem. Phys. 2015, 15, 3063–3075. [Google Scholar] [CrossRef] [Green Version]
  35. Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103, 4605–4638. [Google Scholar] [CrossRef]
  36. Hinks, M.L.; Montoya-Aguilera, J.; Ellison, L.; Lin, P.; Laskin, A.; Laskin, J.; Shiraiwa, M.; Dabdub, D.; Nizkorodov, S.A. Effect of relative humidity on the composition of secondary organic aerosol from the oxidation of toluene. Atmos. Chem. Phys. Online 2018, 18, 1643–1652. [Google Scholar] [CrossRef] [Green Version]
  37. Sato, K.; Takami, A.; Kato, Y.; Seta, T.; Fujitani, Y.; Hikida, T.; Shimono, A.; Imamura, T. AMS and LC/MS analyses of SOA from the photooxidation of benzene and 1,3,5-trimethyl benzene in the presence of NOx: Effects of chemical structure on SOA aging. Atmos. Chem. Phys. 2012, 12, 4667–4682. [Google Scholar] [CrossRef] [Green Version]
  38. Suh, I.; Zhang, R.; Molina, L.T.; Molina, M.J. Oxidation Mechanism of Aromatic Peroxy and Bicyclic Radicals from OH-Toluene Reactions. J. Am. Chem. Soc. 2003, 125, 12655–12665. [Google Scholar] [CrossRef]
  39. Carlton, A.G.; Turpin, B.; Altieri, K.E.; Seitzinger, S.; Reff, A.; Lim, H.-J.; Ervens, B. Atmospheric oxalic acid and SOA production from glyoxal: Results of aqueous photooxidation experiments. Atmos. Environ. 2007, 41, 7588–7602. [Google Scholar] [CrossRef]
  40. Laskin, J.; Laskin, J.; Nizkorodov, S.A. Chemistry of Atmospheric Brown Carbon. Chem. Rev. 2015, 115, 4335–4382. [Google Scholar] [CrossRef] [Green Version]
  41. Updyke, K.M.; Nguyen, T.B.; Nizkorodov, S.A. Formation of brown carbon via reactions of ammonia with secondary organic aerosols from biogenic and anthropogenic precursors. Atmos. Environ. 2012, 63, 22–31. [Google Scholar] [CrossRef]
  42. Kampf, C.J.; Jakob, R.; Hoffmann, T. Identification and characterization of aging products in the glyoxal/ammonium sulfate system-implications for light-absorbing material in atmospheric aerosols. Atmos. Chem. Phys. 2012, 12, 6323–6333. [Google Scholar] [CrossRef] [Green Version]
  43. Lee, A.K.Y.; Zhao, R.; Li, R.; Liggio, J.; Li, S.-M.; Abbatt, J.P.D. Formation of Light Absorbing Organo-Nitrogen Species from Evaporation of Droplets Containing Glyoxal and Ammonium Sulfate. Environ. Sci. Technol. 2013, 47, 12819–12826. [Google Scholar] [CrossRef] [PubMed]
  44. Maxut, A.; Mechakra, H.; Nozière, B.; Fenet, B. Formation mechanisms and yields of small imidazoles from reactions of glyoxal with NH4+in water at neutral pH. Phys. Chem. Chem. Phys. 2015, 17, 20416–20424. [Google Scholar] [CrossRef]
  45. Gao, Y.; Wang, H.; Zhang, X.; Jing, S.; Peng, Y.; Qiao, L.; Zhou, M.; Huang, D.D.; Wang, Q.; Li, X.; et al. Estimating Secondary Organic Aerosol Production from Toluene Photochemistry in a Megacity of China. Environ. Sci. Technol. 2019, 53, 8664–8671. [Google Scholar] [CrossRef]
  46. Liu, Y.; Liggio, J.; Staebler, R.; Li, S.-M. Reactive uptake of ammonia to secondary organic aerosols: Kinetics of organonitrogen formation. Atmos. Chem. Phys. 2015, 15, 13569–13584. [Google Scholar] [CrossRef] [Green Version]
  47. Trainic, M.; Riziq, A.A.; Lavi, A.; Flores, J.M.; Rudich, Y. The optical, physical and chemical properties of the products of glyoxal uptake on ammonium sulfate seed aerosols. Atmos. Chem. Phys. 2011, 11, 9697–9707. [Google Scholar] [CrossRef] [Green Version]
  48. Pósfai, M.; Xu, H.; Anderson, J.R.; Buseck, P.R. Wet and dry sizes of atmospheric aerosol particles: An AFM-TEM Study. Geophys. Res. Lett. 1998, 25, 1907–1910. [Google Scholar] [CrossRef] [Green Version]
  49. Cruz, C.N.; Pandis, S.N. Deliquescence and Hygroscopic Growth of Mixed Inorganic-Organic Atmospheric Aerosol. Environ. Sci. Technol. 2000, 34, 4313–4319. [Google Scholar] [CrossRef]
  50. Meyer, N.K.; Duplissy, J.; Gysel, M.; Metzger, A.; Dommen, J.; Weingartner, E.; Alfarra, M.R.; Fletcher, C.; Good, N.; McFiggans, G.; et al. Analysis of the hygroscopic and volatile properties of ammonium sulphate seeded and un-seeded SOA particles. Atmos. Chem. Phys. 2008, 8, 8629–8659. [Google Scholar] [CrossRef] [Green Version]
  51. Hallquist, M.; Wenger, J.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.; Dommen, J.; Donahue, N.M.; George, C.; Goldstein, A.H.; et al. The formation, properties and impact of secondary organic aerosol: Current and emerging issues. Atmos. Chem. Phys. 2009, 9, 5155–5236. [Google Scholar] [CrossRef] [Green Version]
  52. Lu, Z.; Hao, J.; Takekawa, H.; Hu, L.; Li, J. Effect of high concentrations of inorganic seed aerosols on secondary organic aerosol formation in the m-xylene/NOx photooxidation system. Atmos. Environ. 2009, 43, 897–904. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the aerosol albedometer and chamber.
Figure 1. Schematic diagram of the aerosol albedometer and chamber.
Atmosphere 11 00961 g001
Figure 2. The corrected mass concentration of toluene secondary organic aerosol (SOA) formed in absence of seed and in presence of 300 μg/m3 (NH4)2SO4 seed as a function of reaction time.
Figure 2. The corrected mass concentration of toluene secondary organic aerosol (SOA) formed in absence of seed and in presence of 300 μg/m3 (NH4)2SO4 seed as a function of reaction time.
Atmosphere 11 00961 g002
Figure 3. The particle number concentration, mean diameter, extinction and scattering coefficients, SSA and complex refractive index (CRI) of SOA without (NH4)2SO4 seed as a function of reaction time.
Figure 3. The particle number concentration, mean diameter, extinction and scattering coefficients, SSA and complex refractive index (CRI) of SOA without (NH4)2SO4 seed as a function of reaction time.
Atmosphere 11 00961 g003
Figure 4. The particle number concentration, mean diameter, extinction and scattering coefficients, single-scattering albedo (SSA) and CRI of SOA formed in presence of 300 μg/m3 (NH4)2SO4 seed as a function of reaction time.
Figure 4. The particle number concentration, mean diameter, extinction and scattering coefficients, single-scattering albedo (SSA) and CRI of SOA formed in presence of 300 μg/m3 (NH4)2SO4 seed as a function of reaction time.
Atmosphere 11 00961 g004
Figure 5. The (a) corrected maximum mass concentration (b) SSA (c) n and (d) k of SOA particles measured at different concentrations of (NH4)2SO4 seed.
Figure 5. The (a) corrected maximum mass concentration (b) SSA (c) n and (d) k of SOA particles measured at different concentrations of (NH4)2SO4 seed.
Atmosphere 11 00961 g005
Figure 6. Suggested mechanism for aldehyde and carboxylic productions in photooxidation of toluene.
Figure 6. Suggested mechanism for aldehyde and carboxylic productions in photooxidation of toluene.
Atmosphere 11 00961 g006
Figure 7. Negative mass spectrum of SOA without (NH4)2SO4 seed.
Figure 7. Negative mass spectrum of SOA without (NH4)2SO4 seed.
Atmosphere 11 00961 g007
Figure 8. UV-Vis spectra of the SOA (a) without and (b) in presence of 300 μg/m3 (NH4)2SO4 seed.
Figure 8. UV-Vis spectra of the SOA (a) without and (b) in presence of 300 μg/m3 (NH4)2SO4 seed.
Atmosphere 11 00961 g008
Figure 9. Negative mass spectrum of SOA formed in presence of 300 μg/m3 (NH4)2SO4 seed.
Figure 9. Negative mass spectrum of SOA formed in presence of 300 μg/m3 (NH4)2SO4 seed.
Atmosphere 11 00961 g009
Figure 10. Suggest mechanism for imidazole compounds.
Figure 10. Suggest mechanism for imidazole compounds.
Atmosphere 11 00961 g010

Share and Cite

MDPI and ACS Style

Lu, T.; Huang, M.; Zhao, W.; Hu, C.; Gu, X.; Zhang, W. Influence of Ammonium Sulfate Seed Particle on Optics and Compositions of Toluene Derived Organic Aerosol in Photochemistry. Atmosphere 2020, 11, 961. https://doi.org/10.3390/atmos11090961

AMA Style

Lu T, Huang M, Zhao W, Hu C, Gu X, Zhang W. Influence of Ammonium Sulfate Seed Particle on Optics and Compositions of Toluene Derived Organic Aerosol in Photochemistry. Atmosphere. 2020; 11(9):961. https://doi.org/10.3390/atmos11090961

Chicago/Turabian Style

Lu, Tingting, Mingqiang Huang, Weixiong Zhao, Changjin Hu, Xuejun Gu, and Weijun Zhang. 2020. "Influence of Ammonium Sulfate Seed Particle on Optics and Compositions of Toluene Derived Organic Aerosol in Photochemistry" Atmosphere 11, no. 9: 961. https://doi.org/10.3390/atmos11090961

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop