Next Article in Journal
Effect of Biochar and Straw Application on Nitrous Oxide and Methane Emissions from Eutric Regosols with Different pH in Sichuan Basin: A Mesocosm Study
Next Article in Special Issue
Green Leaf Volatiles in the Atmosphere—Properties, Transformation, and Significance
Previous Article in Journal
Theoretical Chemistry and the Calculation of the Atmospheric State
Previous Article in Special Issue
Structural Characterisation of Dimeric Esters in α-Pinene Secondary Organic Aerosol Using N2 and CO2 Ion Mobility Mass Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Secondary Organic Aerosol Formation from Isoprene: Selected Research, Historic Account and State of the Art

1
Department of Pharmaceutical Sciences, Campus Drie Eiken, University of Antwerp, BE-2610 Antwerp, Belgium
2
Department of Chemistry, Ghent University, Krijgslaan 281, S12, BE-9000 Gent, Belgium
*
Author to whom correspondence should be addressed.
Atmosphere 2021, 12(6), 728; https://doi.org/10.3390/atmos12060728
Submission received: 30 April 2021 / Revised: 31 May 2021 / Accepted: 1 June 2021 / Published: 6 June 2021

Abstract

:
In this review, we cover selected research on secondary organic aerosol (SOA) formation from isoprene, from the beginning of research, about two decades ago, to today. The review begins with the first observations of isoprene SOA markers, i.e., 2-methyltetrols, in ambient fine aerosol and focuses on studies dealing with molecular characterization, speciation, formation mechanisms, and source apportionment. A historic account is given on how research on isoprene SOA has developed. The isoprene SOA system is rather complex, with different pathways being followed in pristine and polluted conditions. For SOA formation from isoprene, acid-catalyzed hydrolysis is necessary, and sulfuric acid enhances SOA by forming additional nonvolatile products such as organosulfates. Certain results reported in early papers have been re-interpreted in the light of recent results; for example, the formation of C5-alkene triols. Attention is given to mass spectrometric and separation techniques, which played a crucial role in molecular characterization. The unambiguous structural characterization of isoprene SOA markers has been achieved, owing to the preparation of reference compounds. Efforts have also been made to use air quality data to estimate the influence of biogenic and pollution aerosol sources. This review examines the use of an organic marker-based method and positive matrix factorization to apportion SOA from different sources, including isoprene SOA.

Graphical Abstract

1. Introduction

In this review, selected research on secondary organic aerosol (SOA) from isoprene (2-methyl-1,3-butadiene, C5H8) will be covered from the beginning of research to today. Unlike SOA research from monoterpenes such as α-pinene, SOA research from isoprene started much later, with the discovery of marker compounds, the 2-methyltetrols, which have retained the C5-isoprene skeleton. By now, it is very well established that isoprene is a precursor for SOA, but there has been a time that it was widely accepted that isoprene only resulted in volatile organic compounds (VOCs) upon photo-oxidation in the atmosphere. It was then not surprising that the discovery of SOA from isoprene met with some criticism in the atmospheric research community. The literature on isoprene SOA research is very large, so it is impossible in this review to deal with most papers on this topic. A search on the web-of-science with ‘isoprene’ and ‘secondary organic aerosol’ as topics led to a number of over 1000, with the paper on the discovery of the 2-methyltetrols in Amazonian aerosols the second most cited one [1]. This review begins with the first observations of isoprene SOA marker compounds in ambient aerosol and mainly focuses on studies dealing with molecular characterization, speciation, formation mechanisms, and source apportionment. In this review, the term ‘marker’ is preferred instead of the term ‘tracer’ because organic compounds in the atmosphere generally do not fulfill the requirement of preservation (e.g., due to reactions decreasing their concentrations in the atmosphere). By definition, an atmospheric tracer is “an entity which preserves its identity as it moves with the air from a known source, where the tracer is created or otherwise introduced into the atmosphere, to a known sink where it is destroyed or removed from the atmosphere” [2].
Pathways leading to SOA formation have been intensively explored during the past two decades. The current status is briefly summarized in Figure 1, Figure 2 and Figure 3. Comprehensive reviews are available where the topics of the formation, mass yield, modelling and measurement of isoprene SOA have been documented [3,4,5].

2. Early Research (2004–2014): First Observations, Laboratory Studies, Field Monitoring Studies, Quantitation, and Mechanistic Insights

2.1. First Observations

The molecular characterization of isoprene SOA marker compounds, i.e., the 2-methyltetrols, occurred at the start of intensive research on SOA formation from isoprene almost two decades ago [1]. These compounds were discovered in fine particulate matter (PM; PM2.5 and PM10, PM with an aerodynamic diameter ≤2.5 μm and 10 μm) that was collected in the Amazon region, where isoprene emissions from broadleaf trees are estimated to be very high [9]. Isoprene is quantitatively the most important of the biogenic VOCs, with an estimated global annual emission of about 400–600 TgC. Terrestrial plant foliage is a source of >90% of atmospheric isoprene and it is generally assumed that tropical forests provide most of the global isoprene budget. For the chemical characterization of the 2-methyltetrols, use was made of capillary gas chromatography/mass spectrometry (GC/MS), electron ionization (EI), and the derivatization of polar compounds into trimethylsilyl derivatives. Structurally informative EI and methane chemical ionization (CI) spectra were obtained and interpreted in detail, but to support the structural proposal and unambiguous identification of the proposed organic compounds, reference compounds were also synthesized [1,10]. A tentative formation mechanism involving the gas-phase photo-oxidation of isoprene with the OH radical was proposed, but later on, this mechanism had to be revised [6,11]. The 2-methyltetrols occur in two diastereoisomeric forms, 2-methylthreitol and 2-methylerythritol in a ratio of about 1:3, which upon trimethylsilylation are separated by capillary GC, and each of them exists in two enantiomeric forms (Figure 1). It was not realized at that time that the 2-methyltetrols detected in PM were likely, to a major extent, to be in the form of sulfate esters [6,12], because sulfate esters are not stable upon trimethylsilylation (an acid-catalyzed procedure) and are converted into trimethylsilyl esters. Following the detection of 2-methyltetrols in ambient fine aerosol, laboratory photo-oxidation experiments with isoprene were conducted under controlled conditions and molecular evidence was provided for SOA formation, as discussed in Section 2.2. In addition, the SOA mass yield from isoprene was re-evaluated in laboratory photo-oxidation smog chamber experiments under different conditions, e.g., [13,14] (for a review, see [3]). Even with a low SOA mass yield, the total formation of SOA from isoprene in the ambient atmosphere may be very high and contribute substantially to global SOA production, given the high emission rates of isoprene in forested regions. The SOA yield from isoprene is strongly acid-dependent and closely related to the particle-phase acidity, e.g., [15] but decreases upon increasing humidity, e.g., [16].

2.2. Laboratory Studies

The first laboratory photo-oxidation experiments were performed at the National Exposure Laboratory of the United States Environmental Protection Agency in Research Triangle Park (NC, USA) by Edward Edney, Tadeusz Kleindienst and coworkers, and the conditions chosen simulated regional atmospheric conditions (i.e., in the presence of NOx (NO + NO2) and SO2) [17]. It was suggested that the formation of the 2-methyltetrols was likely acid-catalyzed, although the precise role of sulfuric acid was not clear at that time. It was already evident from this first laboratory study that the formation of 2-methyltetrols was strongly acid-dependent and enhanced in the presence of sulfuric acid. In addition to the 2-methyltetrols, it was also demonstrated that 2-methylglyceric acid, which has partially retained the isoprene skeleton, was formed under the selected conditions. While other polar compounds related to the 2-methyltetrols could also be detected in ambient PM, i.e., C5-alkene triols [18], it was puzzling why these compounds were absent from the isoprene SOA that was generated in the presence of NOx [17]. Later, it became clear that these compounds are characteristic for low-NOx conditions [19]. However, it was shown recently that they might be formed in part by the degradation of methyltetrol organosulfates [20]. More comprehensive laboratory experiments were started in 2004 by John Seinfeld and coworkers, at the California Institute of Technology (Pasadena, CA, USA). They were in a later stage and are still actively pursued in the group of Jason Surratt at the University of North Carolina at Chapel Hill (USA). Photo-oxidation experiments were performed with isoprene with and without the addition of NO to reflect polluted and clean pristine conditions, and markedly different products were formed [19,21]. In the absence of NOx, evidence was found for the formation of both the 2-methyltetrols and the C5-alkene triols, while in the presence of NOx, isoprene was mainly transformed to 2-methylglyceric acid and a multitude of derivatives, including oligomers (di- and trimers) and nitrate, formate and acetate esters. Analytical techniques that were currently in use at that time played a crucial role: for the detection of the 2-methyltetrols, the C5-alkene triols, and 2-methylglyceric acid and its oligomers, gas chromatography (GC/EI-MS) with prior trimethylsilylation, was performed, while for the detection of derivatives of 2-methylglyceric acid, use was made of liquid chromatography (LC) combined with electrospray ionization MS in the negative ion mode [(–)ESI-MS] or direct injection (–)ESI-MS. It is worth mentioning that at that stage, the detailed mechanisms by which the 2-methyltetrols, the C5-alkene triols and 2-methylglyceric acid were formed were not known. With regard to the structural characterization, the unambiguous identification of the 2-methyltetrols was achieved by organic synthesis of the proposed compounds. This approach was not followed for the C5-alkene triols and these compounds were tentatively characterized using detailed interpretation of the EI and methane CI mass spectra of their trimethylsilyl derivatives [18]. The identification of 2-methylglyceric acid was straightforward because this compound was not new and had been characterized before as a human urinary metabolite. Employing GC/EI-MS with prior trimethylsilylation 2-methylglyceric was indicated with high probability using a MS library search, and later confirmed with a synthesized reference compound [22].

2.3. Field Monitoring Studies

Following the laboratory studies mentioned above, numerous field studies have been conducted where isoprene SOA markers were measured in fine ambient aerosol (PM2.5 or PM10) and where mainly 2-methyltetrols, C5-alkene triols and 2-methylglyceric acid were measured using GC/EI-MS with prior trimethylsilylation. As an example, we select a study conducted in the Brazilian Amazon region, where the emission of isoprene is estimated to be very high [23]. Measurements were performed on aerosols that were collected at a pasture site (Rondônia, Brazil) using a high-volume dichotomous sampler (HVDS) within the framework of the 2002 LBA-SMOCC (Large-Scale Biosphere Atmosphere Experiment in Amazonia—Smoke Aerosols, Clouds, Rainfall, and Climate: Aerosols from Biomass Burning Perturb Global and Regional Climate) campaign. The campaign spanned the late dry season (biomass burning), a transition period, and the onset of the wet season (clean conditions). Figure 4 illustrates the power of GC/EI-MS with prior trimethylsilylation for separating isoprene SOA markers with a total ion chromatogram obtained for PM2.5 collected during the transition period of the 2002 LBA-SMOCC campaign. The concentrations of the 2-methyltetrols and the C5-alkene triols were higher in the dry and transition periods than in the wet period; the isoprene SOA marker compounds showed an average concentration of 250 ng m−3 during the dry period versus 157 ng m−3 during the transition period and 52 ng m−3 during the wet period. The percentage of organic carbon (OC) attributable to the isoprene SOA marker compounds was 0.45% in the dry period, 0.83% in the transition period and 1.55% in the wet period. In that study, 2-methylglyceric acid was not measured because its concentrations were very low compared to those of the selected marker compounds (<1 ng m−3). The ratio between the threo and erythro diastereoisomers of the 2-methyltetrols was on average 0.30 and a statistically significant correlation was found between them (R = 0.94), consistent with their formation through the same aerosol process, i.e., photo-oxidation of isoprene. Again, the detailed mechanism by which the selected isoprene SOA marker compounds were formed was not known at the time of the cited study.
In the early field monitoring studies, only concentrations of isoprene SOA marker compounds were determined, and such studies were conducted at several sites, including South and North America, Europe, Asia, and the Arctic. These sites included tropical and boreal forests but also rural and urban sites that are impacted by pollution. Measurements also often included monoterpene SOA, biomass burning and fungal spore markers, providing information on other aerosol sources. The references cited here are limited to our own early field monitoring studies [23,24,25,26,27]. The atmospheric concentrations of the 2-methyltetrols were found to be the highest during daytime, consistent with the emission of isoprene which is both light- and temperature-dependent [28], as well as with OH radical formation, which is a complex daytime process (for a review, see [29]). The OH radical is the main oxidant leading to isoprene SOA formation, but ozone has also been reported to be a minor contributor, e.g., [30].

2.4. Quantitation of Isoprene SOA Markers

The quantitation of molecular markers of isoprene SOA has most commonly been performed with GC/EI-MS with prior trimethylsilylation and LC/(–)ESI-MS, with the latter technique applied to organosulfates and nitrooxy organosulfates (see Section 3). Reference standards are needed for accurate quantitation but are unfortunately only available for a few marker compounds. A list of synthesized reference compounds for isoprene SOA (e.g., 2-methyltetrols, 2-methylglyceric acid) and other biogenic SOA molecular markers is provided in the review by Nozière and coworkers [5].
For quantitation, use has often also been made of surrogate standards with structural properties very similar to the targeted analytes when no pure reference compounds were available; for example, for assessing the amounts of the 2-methyltetrols and C5-alkene triols, the response factor of erythritol and threitol have been applied, e.g., [23,31]. The GC/EI-MS method with prior trimethylsilylation with N,O-bis(trimethylsilyl)trifluoroacetamide (BSTFA) containing an acidic catalyst (1% trimethylchlorosilane (TMSCl)) has generally been performed off-line in early field monitoring studies on PM that was collected on quartz fiber filters and extracted with a polar solvent (e.g., methanol, acetonitrile), e.g., [23,24,25,26,27]. With respect to the choice of an extraction solvent, acetonitrile is to be preferred above methanol, because of the risk of methylation artefacts in acidic medium, e.g., [32]. An online measurement technique, applied successfully in later studies and allowing the assessment of isoprene SOA, e.g., [33], is thermal desorption aerosol gas chromatography (TAG)/MS, where in situ derivatization using N-methyl-N-(trimethylsilyl)trifluoroacetamide (MSTFA) has been implemented [34].
Selected field studies that deal with source apportionment of PM, OC and SOA, comprising isoprene SOA, are covered in this review. These studies involved source apportionment with an organic marker-based method or positive matrix factorization (PMF) and are discussed in Section 5.

2.5. Mechanistic Insights into the Formation of the 2-methyltetrols, 2-methylglyceric acid, and Derivatives

Gas-phase photo-oxidation studies provided important insights into the formation of the 2-methyltetrols and related SOA marker compounds from isoprene. An important breakthrough was the discovery of dihydroxyepoxides, referred to as IEPOX, in the gas phase upon photo-oxidation of isoprene [11]. Under pristine (low-NOx) conditions, isoprene is rapidly oxidized by the OH radical to hydroxyhydroperoxides, which are further oxidized with OH, resulting in the formation of dihydroxyepoxides (i.e., β-IEPOX and δ-IEPOX). These dihydroxyepoxides were identified as crucial intermediates in the formation of isoprene SOA, including 2-methyltetrols, C5-alkene triols, and sulfate esters of methyltetrols and dimers thereof [6]. The latter study clearly revealed that the formation of isoprene SOA requires acidic conditions, i.e., the presence of sulfuric acid. Additional insights were obtained in laboratory experiments where synthesized isoprene epoxydiols were studied for their potential to form SOA via reactive uptake in the presence of acidic seed aerosol [35]. This study demonstrated that the reactive uptake of isoprene epoxydiols explains the formation of known isoprene SOA tracers found in ambient aerosols, including 2-methyltetrols, C5-alkene triols, dimers, and IEPOX-derived organosulfates. In addition, it provided evidence for an unreported acid-catalyzed intramolecular rearrangement of isoprene epoxydiols to cis- and trans-3-methyl-tetrahydrofuran-3,4-diols in the particle phase. Figure 1 summarizes the formation mechanism of major isoprene SOA marker compounds at low-NOx.
Insights were also obtained into the formation of isoprene SOA that is formed under high-NOx conditions, simulating a polluted environment. It was demonstrated by Sato [7] that SOA from laboratory experiments of isoprene in the presence of NO contained 2-methyltetrols, methylnitrooxybutanetriols, and methyldinitrooxybutane diols, of which the 2-methyltetrols should be regarded as hydrolysis products formed together with nitric acid (Figure 2). The reaction of isoprene with OH radicals proceeds through addition to one of the double bonds. The adduct radical thus produced has two resonance structures as a result of the addition of an OH radical to a terminal carbon; for simplicity, only the subsequent reactions of the former radical are shown. The adduct radicals react with O2 under atmospheric conditions to form peroxy radicals (RO2), which react with NO to form either alkoxy radicals or organic nitrates in the presence of NO. The organic nitrates produced through the reaction RO2 + NO → RO + NO2 during the isoprene oxidation are methylnitrooxybutenol isomers. The subsequent oxidation of methylnitrooxybutenols leads to methyldinitrooxybutanediol isomers (A). If reaction RO2 + R’CH2O2 → ROH + O2 + R’CHO occurs, methylnitrooxybutanetriol isomers (B) are also formed. This RO2 + R’CH2O2 type reaction may occur at very low NO levels observed upon SOA formation, but the concentration of the 2-methyltetrols and the methylnitrooxybutanetriols produced should be lower than that of the nitrooxypolyols in the presence of NOx. Nitrooxypolyols (products A and B) produced by the gas-phase reaction are absorbed on the existing particles following the establishment of a gas–particle equilibrium. Methyldinitrooxy-butanediols (A) hydrolyze to form methylnitrooxybutanetriols (B) and nitric acid, and methylnitrooxy-butanetriols undergo further hydrolysis to form 2-methyltetrols (C) and nitric acid.
With regard to the formation of 2-methylglyceric acid at high-NOx, 2-methyloxirane-2-carboxylic acid (methacrylic acid epoxide; MAE), arising from the decomposition of the OH adduct of methacryloylperoxynitrate, was shown to be a precursor [8,36]. The mechanism proposed in Figure 3 shows the route to methacryloylperoxynitrate (MPAN) via H-abstraction of the branching of methacrolein with OH. The OH adduct of MPAN undergoes an intramolecular rearrangement to MAE, which yields high-NOx SOA compounds, including 2-methylglyceric acid (2-MG), its organosulfate, 2-MG dimer, and an organic nitrate, 2-hydroxy-2-methyl-3-nitrooxypropionic acid.

3. Isoprene SOA-Related Organosulfates: Detection, Structural Characterization, Formation Mechanisms, Analytical Challenges, and Degradation

A comprehensive review on organosulfates in ambient aerosol covering the state of knowledge and future research directions on formation, abundance, fate and importance has recently been published [37]. In the current review, we focus on organosulfate formation from isoprene.

3.1. Detection and Structural Characterization

The application of LC/(–)ESI-MS led to the detection of organosulfates (OSs) related to isoprene and monoterpene SOA in ambient PM [12,38]. Due to the hydrophilicity of organosulfates related to isoprene, C18-reversed-phase LC columns showing increased polar retention had to be employed, such as the Atlantis dC18 and tC18 column (Waters). These columns contain di- or trifunctionally bonded C18 alkyl residues preventing stationary phase collapse when an aqueous mobile phase is used. Both regular high (HPLC) and ultra-high-performance LC (UHPLC) have been employed with the latter technique using columns containing smaller stationary phase particles by which larger pressures can be applied and elution times are shorter. The major organosulfates related to isoprene SOA were identified in ambient PM2.5 as organosulfate esters of methyltetrols (MW 216), but in addition, a number of less abundant organosulfates could be identified, including OSs related to volatile oxidation products of isoprene, i.e., methacrolein (MW 200) and methyl vinyl ketone (MW 184), and MW 212 OSs.
The major MW 200 OS related to methacrolein could be identified as a derivative of 2-methylglyceric acid with a terminal sulfate group, i.e., 2-hydroxy-2-methyl-3-sulfoxypropionic acid [39], whereas the MW 184 OS related to methyl vinyl ketone could be elucidated as 4-hydroxy-3-sulfoxybutan-2-one [40]. Using detailed interpretation of LC/ (–)ESI-MS data, including tandem MS (MS/MS), a structure was proposed in each case and in a following step, the proposed compound was confirmed using organic synthesis and comparison of its MS and chromatographic characteristics. As an example, we mention a specific feature of the MW 200 OS, 2-hydroxy-2-methyl-3-sulfoxypropionic acid, i.e., its terminal sulfate group. This position was indicated by the lack of a [M − H − CH2O] ion in the MS/MS spectrum of the deprotonated molecule [M − H] at m/z 199, which is characteristic for a terminal CH2OH group. In contrast, in the case of the MW 184 OS related to methyl vinyl ketone, 4-hydroxy-3-sulfoxybutan-2-one, a [M − H − CH2O] ion (m/z 153) in the MS/MS spectrum of the deprotonated molecule [M − H] at m/z 183 provided evidence for a terminal CH2OH group.
The identification of the MW 212 OSs represented a very complex structural puzzle that took a long time to resolve. These OSs were already detected in early studies [38,39]; their cyclic lactone structures were proposed [41] but were only recently structurally confirmed [42]. In an effort to produce the targeted MW 212 OSs and to support their structures, sulfate esters of 3,4-dihydroxy-3-methyl-tetrahydro-2-furanone and 3,4-dihydroxy-4-methyl-tetrahydro-2-furanone were prepared by the aqueous-phase reaction of authentic 3-methyl-2-furanone or 4-methyl-2-furanone with the sulfate radical anion [43]. A mixture of six m/z 211 compounds was obtained, which showed the same retention times and MS properties as those observed for ambient PM2.5 and smog chamber laboratory samples. A detailed MS characterization could be achieved for four of the six isomers. Taking into account that the MW 216 methyltetrol-related OSs are formed from isoprene and are abundantly present in ambient PM2.5, it was logical to propose that they could serve as precursors for the MW 212 OSs. They can be explained by OH oxidation of a terminal hydroxyl group to a carboxylic acid group, followed by lactonization [43] or, alternatively, by further reaction of the MW 214 OSs, formed by OH oxidation of a terminal hydroxyl group to an aldehyde group, followed by cyclic hemiacetal formation. The MW 212 OSs can be regarded as processed or aged isoprene SOA marker compounds. During a field monitoring study conducted during a 2007 summer episode in Brasschaat, Belgium [44], it was noted that the MW 212 OSs correlated very well (R = 0.93) with 3-methylbutanetricarboxylic acid (MBTCA), a known marker compound for aged α-pinene SOA [45]. MW 214 methyltetrol-related OSs have also been reported, of which four of the five isomers were identified as cyclic hemiacetals and can be regarded as intermediates in the formation of the MW 212 OSs [41,46]. A recent study provided evidence that 2-methyltetrol OSs undergo oxidative aging to OSs with MWs 156, 168, 200, 212, 214, 230 and 232, with some of them also being detected in ambient PM2.5 [47]. It is noted that MW 156, 168 and 200 OSs present in aged summer PM2.5 from K-puszta, Hungary, already received ample attention in early reports focusing on structural characterization by MS/MS [39,40]. For example, the MW 156 and 168 OSs were identified as glycolic and lactic acid OSs, while minor MW 200 OSs were tentatively assigned to diastereoisomers of 2,3-dihydroxybutanoic acid with the sulfate group located at the 2-position (3-hydroxy-2-sulfoxybutanoic acid). It is also worth mentioning that glycolic and lactic acid OSs have been quantified in ambient PM and that it has been proposed that glycolic acid sulfate represents an important marker for atmospheric processes that form organosulfates in ambient particulate matter, at a time that its origin was still elusive [48].
In addition to isoprene SOA-related OSs, isoprene SOA-related nitrooxy organosulfates (NOSs) have been reported in smog chamber experiments performed in the presence of NOx and in ambient PM [38,39,49]. Several derivatives of 2-methyltetrols containing nitrate and sulfate groups have been detected and tentatively structurally characterized (MWs 261, 306, 245, 334 and 332). These markers exist as a mixture of multiple isomers and are partially resolved by C18-reversed-phase LC. It was suggested that the formation of these nitrooxy organosulfates involves the nitration of isoprene in the gas phase, followed by sulfation in the particle phase [38]. In this respect, it was shown that the irradiation of isoprene in the presence of NOx results in SOA containing methylnitrooxybutanetriols and methyldinitrooxybutane diols, e.g., [7,50]. Taking into account the instability of both the nitrate and sulfate group, NOSs are converted to the trimethylsilylated derivatives of their free hydroxyl forms upon GC/EI-MS analysis with prior trimethylsilylation. Using this analysis method, unknown N-containing SOA markers have been reported in smog chamber experiments where isoprene was irradiated in the presence of NOx and ammonium sulfate/sulfuric acid inorganic seed aerosol, which were also observed in ambient PM2.5 [51]. For a comprehensive list of OSs and NOSs characterized in isoprene SOA and ambient PM, see [49].

3.2. Formation Mechanisms

With regard to the formation mechanism of isoprene-related OSs, several mechanisms have been proposed (for a review, see [37]). In addition to the heterogeneous reaction of epoxides with the sulfate anion, as in the case of the OSs related to methyltetrols [6] and 2-methylglyceric acid [8], evidence has been provided for an alternative formation mechanism, i.e., the reaction of alkenes with the sulfate radical anion in aqueous solution [43]. It was shown that the aqueous phase reaction of isoprene, methacrolein and methyl vinyl ketone with the sulfate radical anion results in OSs from methacrolein and methyl vinyl ketone and is a reasonable pathway for their formation [52,53,54,55]. The reaction with the sulfate radical anion has also been exploited in the synthesis of reference compounds, including the MW 212 isoprene-derived OSs [42], and OSs related to green leaf volatiles [56,57].

3.3. Analytical Challenges and Quantitation

The chromatographic analysis of OSs in a complex matrix such as PM extracts is challenging, because of their isomeric complexity, various precursor origins, and their hydrophilicity. Hydrophilic OSs are not only formed from isoprene and their gas-phase oxidation products, methacrolein and methyl vinyl ketone, but also from 2-methyl-3-buten-2-ol [58] and plant leaf volatiles, including 2-E-pentenal, 2-Z-hexenal, 3-Z-hexenal [56,57] and cis-3-hexen-1-ol [59]. In the first studies employing LC, use was mainly made of C18 reversed-phase columns with polar retention, which contain di- or trifunctionally bonded C18 alkyl residues that prevent stationary phase collapse when an aqueous mobile phase is used. In later studies, the excellent retention of hydrophilic isoprene-related OSs could be achieved with hydrophilic interaction liquid chromatography (HILIC) on an amide stationary phase [20,46,60,61]. Figure 5 compares the analytical power of HPLC on a reversed phase C18 column (i.e., Waters Acquity HSS T3) and HILIC on an amide stationary phase (i.e., Waters Acquity BEH Amide) for separating MW 216 OSs derived from isoprene [46]. Two peaks are noted upon C18 reversed-phase HPLC, corresponding with previous reports [38,39], while four peaks are detected upon HILIC on an amide column, also consistent with published data [41]. In a more recent report [20], the characterization of methyltetrol organosulfates originating from the sulfation of β- and δ-IEPOX could be obtained by which their six possible isomeric forms were separated, i.e., three positional isomers, each occurring as diastereoisomeric pairs. In addition, the developed HILIC method enabled their simultaneous analysis with 2-methyltetrols, the corresponding free hydroxyl forms. It was demonstrated that conventional GC/EI-MS analyses overestimate 2-methyltetrols by up to 188%, resulting (in part) from the degradation of methyltetrol sulfates. In addition, C5-alkene triols and 3-methyltetrahydrofuran-3,4-diols were not detected with HILIC/ESI-MS and were explained by the degradation of 2-methyltetrol sulfates and 3-methyltetrol sulfates, respectively.
For quantitation, reference standards recently became available for isoprene SOA-related organosulfates. Methyltetrol sulfates (MW 216) have been synthesized following different sulfation strategies starting from δ-IEPOX or 2-methyltetrol [20,62]. They were employed in the study by Glasius and coworkers [63] to determine concentrations of methyltetrol OSs as well as other related OSs (MWs 156, 170, 184, 200, 212 and 214) in Amazonian PM1. In addition, MW 212 OSs have been synthesized starting from 3-methyl-2-furanone or 4-methyl-2-furanone with sulfate radical anions [42]. Procedures for the synthesis of low MW OSs associated with the aging of methyltetrol OSs, i.e., glycolic acid and lactic acid sulfate (MW 156 and 168), have also been reported [47,48]. Quantitative determinations have been performed in several studies but in most cases should be regarded as semi-quantitative because of several uncertainties, including the lack of reference compounds, e.g., [32,63,64,65]. In the study by Gómez-González and coworkers [44], where isoprene SOA-related OSs (MWs of 216 and 212) and NOSs (MWs 261 and 306) were measured, ethanesulfonic acid was used as a surrogate standard, because of its close LC retention time and acidic character, while sebacic acid was chosen as an internal recovery standard to compensate for losses during the sample work-up procedure. In the study by Kristensen and Glasius [32], camphorsulfonic acid was used as a surrogate standard to determine concentrations of MW 216 methyltetrol OSs. Using this method, it could be shown that the absolute concentrations of MW 216 OSs were underestimated 5–10 times compared to determinations where OSs derived from the dihydroxy epoxide of butadiene, structurally related to the MW 216 OSs, were employed [65]. Nevertheless, semi-quantitative methods for the determination of organosulfates can provide important information on variation in levels and time trends and remain useful for marker-based source apportionment.
Direct MS analysis (without chromatography), i.e., nanoelectrospray (nanoESI) ultra-high-resolution mass spectrometry (UHRMS), has potential to obtain valuable insights into the enormous complexity of biogenic and anthropogenic aerosol, e.g., [66,67]. For example, in the last cited study, this approach was applied to the molecular characterization of ambient PM2.5 collected during dry and wet seasons at a site in central Amazonia. Comprehensive mass spectral data evaluation methods (i.e., Kendrick mass defect, Van Krevelen diagrams, carbon oxidation state and aromaticity equivalent) were used to identify compound classes and mass distributions of the detected species. Isoprene-derived organosulfate (IEPOX OS) was found to correspond to the most abundant ion in most of the analyzed samples and strongly followed the concentration trends of the gas-phase anthropogenic tracers, confirming its mixed anthropogenic–biogenic origin.

3.4. Degradation of Organosulfates

With regard to the possible degradation of organosulfates, it has been reported that trimethylsilylated sulfuric acid is formed upon the trimethylsilylation of isoprene SOA generated in the presence of SO2 and analyzed by GC/MS [51]. Trimethylsilylation is commonly performed with a trimethylsilylation reagent (i.e., BSTFA or MSTFA) in the presence of an acidic catalyst (i.e., trimethylsilylchlorosilane). The formation of the 2-methyltetrols and related isoprene SOA markers (i.e., C5-alkene triols, 2-methylglyceric acid) is thus expected when PM extracts containing organosulfates of the 2-methyltetrols, 2-methylglyceric acid, and corresponding oligomers or derivatives, are trimethylsilylated. Moreover, the acid-catalyzed degradation can be enhanced by the release of sulfuric acid (and/or nitric acid). Nevertheless, the 2-methyltetrols and C5-alkene triols formed by the degradation of OSs upon acid-catalyzed trimethylsilylation can serve as a proxy for isoprene SOA and remain useful in marker-based source apportionment.

4. Markers for Isoprene SOA Aging

In recent studies, attention has been given to the structural characterization of aged or processed highly oxygenated isoprene SOA markers. In Section 3.1, we already discussed MW 212 organosulfates, which can be considered as markers for aged isoprene SOA. In addition, seven hydroxy acids were identified using GC/MS with prior derivatization as highly oxygenated markers for aged isoprene SOA, including 2-methyltartronic acid, 2-methylerythronic acid, 3-methylerythronic acid, 2-methylthreonic acid, 3-methylthreonic acid, erythro-methyltartaric acid, and threo-methyltartaric acid [68]. Reference compounds were synthesized to support the proposed structures. Of the identified markers, methyltartaric acids and 2- and 3-methylthreonic acids were found to be representative of aged isoprene aerosol because they occur both in laboratory chamber aerosol and ambient PM2.5 from different locations in the USA and Poland. It is worth noting that diastereoisomeric C5-dihydroxypentanedioic acids (MW 164) were already reported in Amazonian aerosol and mentioned in the supplement of the early paper by Claeys and coworkers [1] on the discovery of the 2-methyltetrols. The MW 164 compounds were tentatively assigned in that paper to linear C5-dihydroxypentanedioic acids but taking into account the EI-MS data and the recent paper by Jaoui and coworkers [68], their structures should be revised as threo- and erythro-methyltartaric acids (2,3-dihydroxy-2-methylbutanedioic acids). Once again, it turned out that structures of unknown compounds can only be identified unambiguously with synthesized reference compounds. Interestingly, organosulfates related to 2-methyltartaric acid (MW 244) and 2-methylthreonic acid (MW 230) have also been identified in smog chamber aerosol, generated in the presence of NOx and ammonium sulfate/sulfuric acid inorganic seed aerosol, and in ambient PM, using the detailed interpretation of ESI mass spectral data [49]. As discussed above, they likely undergo acid-catalyzed degradation in the GC/EI-MS analysis with prior trimethylsilylation, resulting in the TMS-derivatives of the free hydroxyl forms.

5. Source Apportionment

Two approaches for source apportionment are covered in this section, i.e., an organic marker-based method and positive matrix factorization, where organic markers for isoprene SOA were measured in PM1, PM2.5 or PM10, thereby mainly using capillary GC/EI-MS with prior trimethylsilylation and in one case HILIC/ESI-MS for the determination of isoprene SOA-related organosulfates.

5.1. The Organic Marker-Based Approach

An organic marker-based method containing laboratory and field study components was developed by Kleindienst and coworkers [69]. In it, the mass concentrations of organic markers are multiplied by experimentally determined conversion factors to obtain the organic mass that is attributed to the sources and thus to estimate the SOA contributions of biogenic and anthropogenic hydrocarbons to ambient OC concentrations in PM2.5, or thus the secondary OC (SOC). An advantage of this approach is that it can be applied separately on each individual sample and thus also on a single sample. Kleindienst and coworkers [69] performed smog chamber experiments in the laboratory whereby isoprene, α-pinene, β-caryophyllene, and toluene were individually irradiated in the presence of NOx. The method was first applied to 33 PM2.5 samples collected at Research Triangle Park (NC, USA), located in southeastern USA and impacted by urban pollution. The samples were analyzed for (1) molecular markers observed in the laboratory irradiations, (2) levoglucosan, a biomass burning marker, and (3) total OC. For each of the four hydrocarbons, the SOA contributions to ambient OC concentrations were estimated using the organic marker concentrations and the laboratory-derived mass fractions. The estimates show SOA formation from isoprene, α-pinene, β-caryophyllene, and toluene contributed significantly to the ambient OC concentrations. The relative contributions were highly seasonal, with biomass burning in the winter accounting for more than 50% of the OC concentrations, while SOA contributions remained low. However, during the 6-month period between May and October, SOA from the precursor hydrocarbons contributed more than 40% of the measured OC concentration. It was pointed out that the organic marker-based method is subject to considerable uncertainty due to the simplification of replacing the complex set of chemical reactions responsible for SOA with a laboratory-derived single-valued mass fraction. The results suggest that this approach is useful to identify major sources of SOA. Following the development of this method, it has been applied in several field campaign studies that were conducted at locations where biogenic and anthropogenic SOA formation is believed to play a role. In order to obtain a more complete source apportionment, including primary organic aerosol (POA), measurements often also included key molecular markers for anthropogenic and biogenic POA, allowing for the estimation of the contribution of biomass burning, fungal spores, plant pollen, plant detritus, coal burning, meat cooking, waste burning, vehicle exhaust, et cetera, e.g., [69,70,71,72,73,74,75,76,77,78,79,80].

5.2. Positive Matrix Factorization

An approach that has been applied more recently in field campaign studies for source apportionment of PM, OC and SOA, and where key molecular marker compounds for different SOA and POA sources can be included, is positive matrix factorization, a mathematical receptor model that provides scientific support for the development and review of air quality standards (for a review, see [81]). The fundamental principle of source/receptor relationships is that mass conservation can be assumed, and a mass balance analysis can be used to identify and apportion sources of airborne PM in the atmosphere. The approach to obtaining a data set for receptor modeling is to determine a large number of chemical constituents such as molecular markers for SOA and POA, trace elements, inorganic ions, and light absorption data. A large sample set is needed in PMF, preferably consisting of at least 100 samples. It could be demonstrated in several studies that PMF including molecular markers for biogenic and anthropogenic SOA has a distinct advantage in resolving isoprene SOA and other SOA, and in capturing aged SOA, e.g., [82,83,84,85,86]. Generally, a good agreement was found between the PMF-derived results with estimated SOA concentrations based on SOA marker yield estimates obtained from smog chamber experiments, e.g., [81,85], except in cases where there was evidence for aging, e.g., [84]. In a recent study, organosulfates have been included in the PMF data base, allowing additional insights into sources of biogenic SOC, such as isoprene SOC formed under low-NOx conditions, isoprene SOC formed under high-NOx conditions, SOC formed by photochemical reactions, oxidatively aged biogenic SOC and sulfuric acid-influenced SOC [87]. It should be noted that the organic marker-based approach and PMF can be used on the same data set and that the results of both can be compared to estimate the uncertainties that are associated with each of the two approaches, as was performed in [85].
With regard to source apportionment of OC and PM where isoprene SOA is resolved, it is noted that the PMF method has also been applied successfully on data bases containing data other than GC/EI-MS and/or LC/MS measurements of isoprene SOA markers compounds. This approach involves low- and high-resolution Aerodyne Mass Spectrometer (AMS) measurements, where a highly characteristic IEPOX SOA-related ion signal is generated at m/z 82 (C5H6O+), e.g., [88,89,90]. Using PMF, an IEPOX-SOA factor could be extracted, which strongly correlated with 2-methyltetrols, C5-alkene triols, IEPOX-derived organosulfates, and dimers of organosulfates, confirming the role of IEPOX chemistry as the source [91].

6. Overview of Molecular Markers

Table 1 summarizes the molecular markers for isoprene SOA that were discussed in this review and are useful for source apportionment. Molecular structures are presented, and information is given about the chemical name, MW, formation mechanism (figure), analysis method, and selected references.

7. Concluding Remarks

Understanding the formation of isoprene SOA and assessing its relevance for air quality has been a process of advancing insights, where analytical developments also played an important role. In this review, we have presented a historical account of isoprene SOA research which started with the field detection of the 2-methyltetrols in Amazonian aerosols. For SOA formation from isoprene, acid-catalyzed hydrolysis is necessary, and sulfuric acid enhances SOA by forming additional nonvolatile products such as organosulfates. The isoprene SOA system is very complex since it involves the reaction of gas-phase oxidation products, photo-oxidation at low- and high-NOx, nitration reactions in the gas phase, sulfation reactions in the particle phase, and aging. In early field monitoring campaigns, marker compounds for isoprene SOA were measured using capillary GC/EI-MS with prior trimethylsilylation, which enabled the detection of the 2-methyltetrols as major products and the separation of their diastereoisomeric forms, 2-methylthreitol and 2-methylerythritol, as well of 2-methylglyceric acid. Advanced analytical techniques based on LC/MS have become available, with HILIC being a useful method for separating polar isoprene SOA-related organosulfates that exist in different positional isomeric and diastereoisomeric forms. The structural characterization of isoprene SOA markers has been a tremendous challenge, and it has become evident that synthesized reference compounds are needed for their unambiguous characterization. Multifunctional isoprene SOA markers containing nitrooxy, sulfoxy and hydroxyl groups exist as complex isomeric mixtures and warrant further characterization. There is also still limited insight into the formation of aged isoprene SOA marker compounds, such as the MW 212 and 214 organosulfates. Evidence has been obtained that methyltetrol organosulfates undergo degradation upon GC/EI-MS analysis with prior trimethylsilylation to 2-methyltetrols and C5-alkene triols. Source apportionment of SOA, OC and PM, incorporating isoprene SOA employing a marker-based approach or PMF has been developed. It has been shown that for the latter approach, molecular markers are needed to resolve isoprene SOA. In most cases, isoprene SOA markers have been determined with GC/EI-MS with prior trimethylsilylation and only in a selected case markers of the OS type, determined with HILIC/(–)ESI-MS, have been included in PMF. More refined source apportionment allowing for the assessment of aged isoprene SOA is to be expected in the future when suitable markers are included in the analysis.
During the past two decades, detailed insights have been obtained into the fate of isoprene in the atmosphere, and over the years, we have seen that isoprene science is continuously progressing and has not revealed its secrets all at once.

Author Contributions

M.C.: writing; W.M.: contribution to Section 5 (source apportionment) and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Claeys, M.; Graham, B.; Vas, G.; Wang, W.; Vermeylen, R.; Pashynska, V.; Cafmeyer, J.; Guyon, P.; Andreae, M.O.; Artaxo, P.; et al. Formation of secondary organic aerosols through photooxidation of isoprene. Science 2004, 303, 1173. [Google Scholar] [CrossRef] [Green Version]
  2. Sheppard, P.A. Atmospheric tracers and the study of the general circulation of the atmosphere. Rep. Prog. Phys. 1963, 26, 213. [Google Scholar] [CrossRef]
  3. Carlton, A.G.; Wiedinmyer, C.; Kroll, J.H. A review of secondary organic aerosol (SOA) formation from isoprene. Atmos. Chem. Phys. 2009, 9, 4987. [Google Scholar] [CrossRef] [Green Version]
  4. Hallquist, M.; Wenger, J.C.; 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. [Google Scholar] [CrossRef] [Green Version]
  5. Nozière, B.; Kalberer, M.; Claeys, M.; Allan, J.; D’Anna, B.; Decesari, S.; Finessi, E.; Glasius, M.; Grgic, I.; Hamilton, J.F.; et al. The molecular identification of organic compounds in the atmosphere: State of the art and challenges. Chem. Rev. 2015, 115, 3919. [Google Scholar] [CrossRef] [PubMed]
  6. Surratt, J.D.; Chan, A.W.H.; Eddingsaas, N.C.; Chan, M.N.; Loza, C.L.; Kwan, A.J.; Hersey, S.P.; Flagan, R.C.; Wennberg, P.O.; Seinfeld, J.H. Reactive intermediates revealed in secondary organic aerosol formation from isoprene. Proc. Natl. Acad. Sci. USA 2010, 107, 6640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Sato, K. Detection of nitrooxypolyols in secondary organic aerosol formed from the oxidation of conjugated dienes under high-NOx conditions. Atmos. Environ. 2008, 42, 6851. [Google Scholar] [CrossRef]
  8. Lin, Y.H.; Zhang, H.F.; Pye, H.O.T.; Zhang, Z.F.; Marth, W.J.; Park, S.; Arashiro, M.; Cui, T.Q.; Budisulistiorini, H.; Sexton, K.G.; et al. Epoxide as a precursor to secondary organic aerosol formation from isoprene photooxidation in the presence of nitrogen oxides. Proc. Natl. Acad. Sci. USA 2013, 110, 6718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Guenther, A.; Karl, T.; Harley, P.; Wiedinmyer, C.; Palmer, P.I.; Geron, C. Estimates of global terrestrial isoprene emissions using MEGAN (Model of Emissions of Gases and Aerosols from Nature). Atmos. Chem. Phys. 2006, 6, 3181. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, W.; Vas, G.; Dommisse, R.; Loones, K.; Claeys, M. Fragmentation study of diastereoisomeric 2-methyltetrols, oxidation products of isoprene, as their trimethylsilyl ethers, using gas chromatography/ion trap mass spectrometry. Rapid Commun. Mass Spectrom. 2004, 18, 1787. [Google Scholar] [CrossRef]
  11. Paulot, F.; Crounse, J.D.; Kjaergaard, H.G.; Kürten, A.; Clair, J.M.; Seinfeld, J.H.; Wennberg, P.O. Unexpected epoxide formation in the gas-phase photooxidation of isoprene. Science 2009, 325, 730. [Google Scholar] [CrossRef] [Green Version]
  12. Surratt, J.D.; Kroll, J.H.; Kleindienst, T.E.; Edney, E.O.; Claeys, M.; Sorooshian, A.; Ng, N.L.; Offenberg, J.H.; Lewandowski, M.; Jaoui, M.; et al. Evidence for organosulfates in secondary organic aerosol. Environ. Sci. Technol. 2007, 41, 517. [Google Scholar] [CrossRef]
  13. Kroll, J.H.; Ng, N.L.; Murphy, S.M.; Flagan, R.C.; Seinfeld, J.H. Secondary organic aerosol formation from isoprene photooxidation under high-NOx conditions. Geophys. Res. Lett. 2005, 32, L18808. [Google Scholar] [CrossRef] [Green Version]
  14. Kroll, J.H.; Ng, N.L.; Murphy, S.M.; Flagan, R.C.; Seinfeld, J.H. Secondary organic aerosol formation from isoprene photooxidation. Environ. Sci. Technol. 2006, 40, 1869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Surratt, J.D.; Lewandowski, M.; Offenberg, J.H.; Jaoui, M.; Kleindienst, T.E.; Edney, E.O.; Seinfeld, J.H. Effect of acidity on secondary organic aerosol formation from isoprene. Environ. Sci. Technol. 2007, 41, 5363. [Google Scholar] [CrossRef] [PubMed]
  16. Lewandowski, M.; Jaoui, M.; Offenberg, J.H.; Krug, J.D.; Kleindienst, T.E. Atmospheric oxidation of isoprene and 1,3-butadiene: Influence of aerosol acidity and relative humidity on secondary organic aerosol. Atmos. Chem. Phys. 2015, 15, 3773. [Google Scholar] [CrossRef] [Green Version]
  17. Edney, E.O.; Kleindienst, T.E.; Jaoui, M.; Lewandowski, M.; Offenberg, J.H.; Wang, W.; Claeys, M. Formation of 2-methyl tetrols and 2-methylglyceric acid in secondary organic aerosol from laboratory irradiated isoprene/NOx/SO2/air mixtures and their detection in ambient PM2.5 samples collected in the Eastern United States. Atmos. Environ. 2005, 39, 5281. [Google Scholar] [CrossRef]
  18. Wang, W.; Kourtchev, I.; Graham, B.; Cafmeyer, J.; Maenhaut, W.; Claeys, M. Characterization of oxygenated derivatives of isoprene related to 2-methyltetrols in Amazonian aerosols using trimethylsilylation and gas chromatography/ion trap mass spectrometry. Rapid Commun. Mass Spectrom. 2005, 19, 1343. [Google Scholar] [CrossRef]
  19. Surratt, J.D.; Murphy, S.M.; Kroll, J.H.; Ng, N.L.; Hildebrandt, L.; Sorooshian, A.; Szmigielski, R.; Vermeylen, R.; Maenhaut, W.; Claeys, M.; et al. Chemical composition of secondary organic aerosol formed from the photooxidation of isoprene. J. Phys. Chem. A 2006, 110, 9665. [Google Scholar] [CrossRef] [Green Version]
  20. Cui, T.Q.; Zeng, Z.X.; dos Santos, E.O.; Zhang, Z.F.; Chen, Y.Z.; Zhang, Y.; Rose, C.A.; Budisulistiorini, S.H.; Collins, L.B.; Bodnar, W.M.; et al. Development of a hydrophilic interaction liquid chromatography (HILIC) method for the chemical characterization of water-soluble isoprene epoxydiol (IEPOX)-derived secondary organic aerosol. Environ. Sci. Process. Impacts 2018, 20, 1524. [Google Scholar] [CrossRef]
  21. Szmigielski, R.; Surratt, J.D.; Vermeylen, R.; Szmigielska, K.; Kroll, J.H.; Ng, N.L.; Murphy, S.M.; Sorooshian, A.; Seinfeld, J.H.; Claeys, M. Characterization of 2-methylglyceric acid oligomers in secondary organic aerosol formed from the photooxidation of isoprene using trimethylsilylation and gas chromatography/ion trap mass spectrometry. J. Mass Spectrom. 2007, 42, 101. [Google Scholar] [CrossRef]
  22. Claeys, M.; Wang, W.; Ion, A.C.; Kourtchev, I.; Gelencsér, A.; Maenhaut, W. Formation of secondary organic aerosols from isoprene and its gas-phase oxidation products through reaction with hydrogen peroxide. Atmos. Env. 2004, 38, 4093. [Google Scholar] [CrossRef]
  23. Claeys, M.; Kourtchev, I.; Pashynska, V.; Vas, G.; Vermeylen, R.; Wang, W.; Cafmeyer, J.; Chi, X.; Artaxo, P.; Andreae, M.O.; et al. Polar organic marker compounds in atmospheric aerosols during the LBA-SMOCC 2002 biomass burning experiment in Rondônia, Brazil: Sources and source processes, time series, diel variations and size distributions. Atmos. Chem. Phys. 2010, 10, 9319. [Google Scholar] [CrossRef] [Green Version]
  24. Ion, A.C.; Vermeylen, R.; Kourtchev, I.; Cafmeyer, J.; Chi, X.; Gelencsér, A.; Maenhaut, W.; Claeys, M. Polar organic compounds in rural PM2.5 aerosols from K-puszta, Hungary, during a 2003 summer field campaign: Sources and diel variations. Atmos. Chem. Phys. 2005, 5, 1805. [Google Scholar] [CrossRef] [Green Version]
  25. Kourtchev, I.; Ruuskanen, T.; Maenhaut, W.; Kulmala, M.; Claeys, M. Observation of 2-methyltetrols and related photo-oxidation products of isoprene in boreal forest aerosols from Hyytiälä, Finland. Atmos. Chem. Phys. 2005, 5, 2761. [Google Scholar] [CrossRef] [Green Version]
  26. Kourtchev, I.; Warnke, J.; Maenhaut, W.; Hoffmann, T.; Claeys, M. Polar organic marker compounds in PM2.5 aerosol from a mixed forest site in Western Germany. Chemosphere 2008, 73, 1308. [Google Scholar] [CrossRef]
  27. Kourtchev, I.; Ruuskanen, T.M.; Keronen, P.; Sogacheva, L.; Dal Maso, M.; Reissell, A.; Chi, X.; Vermeylen, R.; Kulmala, M.; Maenhaut, W.; et al. Determination of isoprene and α-/β-pinene oxidation products in boreal forest aerosols from Hyytiälä, Finland: Diel variations and possible link with particle formation events. Plant Biol. 2008, 10, 138. [Google Scholar] [CrossRef] [PubMed]
  28. Sharkey, T.D.; Yeh, S.S. Isoprene emission from plants. Annu. Rev. Plant Phys. 2001, 52, 407. [Google Scholar] [CrossRef]
  29. Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental implications of hydroxyl radicals (•OH). Chem. Rev. 2015, 115, 13051. [Google Scholar] [CrossRef]
  30. Kleindienst, T.E.; Lewandowski, M.; Offenberg, J.H.; Jaoui, M.; Edney, E.O. Ozone-isoprene reaction: Re-examination of the formation of secondary organic aerosol. Geophys. Res. Lett. 2007, 34, L01805. [Google Scholar] [CrossRef]
  31. Clements, A.L.; Seinfeld, J.H. Detection and quantification of 2-methyltetrols in ambient aerosol in the Southeastern United States. Atmos. Environ. 2007, 41, 1825. [Google Scholar] [CrossRef]
  32. Kristensen, K.; Glasius, M. Organosulfates and oxidation products from biogenic hydrocarbons in fine aerosols from a forest in North West Europe during spring. Atmos. Environ. 2011, 45, 4546. [Google Scholar] [CrossRef]
  33. Yee, L.D.; Isaacman-VanWertz, G.; Wernis, R.A.; Kreisberg, N.M.; Glasius, M.; Riva, M.; Surratt, J.D.; de Sa, S.S.; Martin, S.T.; Alexander, M.L.; et al. Natural and anthropogenically influenced isoprene oxidation in Southeastern United States and Central Amazon. Environ. Sci. Technol. 2020, 54, 5980. [Google Scholar] [CrossRef] [PubMed]
  34. Isaacman, G.; Kreisberg, N.M.; Yee, L.D.; Worton, D.R.; Chan, A.W.H.; Moss, J.A.; Hering, S.V.; Goldstein, A.H. Online derivatization for hourly measurements of gas- and particle-phase semi-volatile oxygenated organic compounds by thermal desorption aerosol gas chromatography (SV-TAG). Atmos. Meas. Tech. 2014, 7, 4417. [Google Scholar] [CrossRef] [Green Version]
  35. Lin, Y.H.; Zhang, Z.F.; Docherty, K.S.; Zhang, H.F.; Budisulistiorini, S.H.; Rubitschun, C.L.; Shaw, S.L.; Knipping, E.M.; Edgerton, E.S.; Kleindienst, T.E.; et al. Isoprene epoxydiols as precursors to secondary organic aerosol formation: Acid-catalyzed reactive uptake studies with authentic compounds. Environ. Sci. Technol. 2012, 46, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhang, H.F.; Lin, Y.H.; Zhang, Z.F.; Zhang, X.L.; Shaw, S.L.; Knipping, E.M.; Weber, R.J.; Gold, A.; Kamens, R.M.; Surratt, J.D. Secondary organic aerosol formation from methacrolein photooxidation: Roles of NOx level, relative humidity and aerosol acidity. Environ. Chem. 2012, 9, 247. [Google Scholar] [CrossRef] [Green Version]
  37. Bruggemann, M.; Xu, R.S.; Tilgner, A.; Kwong, K.C.; Mutzel, A.; Poon, H.Y.; Otto, T.; Schaefer, T.; Poulain, L.; Chan, M.N.; et al. Organosulfates in ambient aerosol: State of knowledge and future research directions on formation, abundance, fate, and importance. Environ. Sci. Technol. 2020, 54, 3782. [Google Scholar] [CrossRef]
  38. Surratt, J.D.; Gómez-González, Y.; Chan, A.W.H.; Vermeylen, R.; Shahgholi, M.; Kleindienst, T.E.; Edney, E.O.; Offenberg, J.H.; Lewandowski, M.; Jaoui, M.; et al. Organosulfate formation in biogenic secondary organic aerosol. J. Phys. Chem. A 2008, 112, 8345. [Google Scholar] [CrossRef] [Green Version]
  39. Gómez-González, Y.; Surratt, J.D.; Cuyckens, F. Szmigielski, R.; Vermeylen, R.; Jaoui, M.; Lewandowski, M.; Offenberg, J.H.; Kleindienst, T.E.; Edney, E.O.; et al. Characterization of organosulfates from the photooxidation of isoprene and unsaturated fatty acids in ambient aerosol using liquid chromatography/(–) electrospray ionization mass spectrometry. J. Mass Spectrom. 2008, 43, 371. [Google Scholar] [CrossRef]
  40. Safi Shalamzari, M.; Ryabtsova, O.; Kahnt, A.; Vermeylen, R.; Herent, M.F.; Quetin-Leclercq, J.; Van der Veken, P.; Maenhaut, W.; Claeys, M. Mass spectrometric characterization of organosulfates related to secondary organic aerosol from isoprene. Rapid Commun. Mass Spectrom. 2013, 27, 784. [Google Scholar] [CrossRef]
  41. Hettiyadura, A.P.S.; Stone, E.A.; Kundu, S.; Baker, Z.; Geddes, E.; Richards, K.; Humphry, T. Determination of atmospheric organosulfates using HILIC chromatography with MS detection. Atmos. Meas. Tech. 2015, 8, 2347. [Google Scholar] [CrossRef] [Green Version]
  42. Wach, P.; Spólnik, G.; Surratt, J.D.; Blaziak, K.; Rudzinski, K.J.; Lin, Y.H.; Maenhaut, W.; Danikiewicz, W.; Claeys, M.; Szmigielski, R. Structural characterization of lactone-containing MW 212 organosulfates originating from isoprene oxidation in ambient fine aerosol. Environ. Sci. Technol. 2020, 54, 1415. [Google Scholar] [CrossRef] [PubMed]
  43. Rudziński, K.J.; Gmachowski, L.; Kuznietsova, I. Reactions of isoprene and sulphoxy radical-anions—A possible source of atmospheric organosulphites and organosulphates. Atmos. Chem. Phys. 2009, 9, 2129. [Google Scholar] [CrossRef] [Green Version]
  44. Gómez-González, Y.; Wang, W.; Vermeylen, R.; Chi, X.; Neirynck, J.; Janssens, I.A.; Maenhaut, W.; Claeys, M. Chemical characterisation of atmospheric aerosols during a 2007 summer field campaign at Brasschaat, Belgium: Sources and source processes of biogenic secondary organic aerosol. Atmos. Chem. Phys. 2012, 12, 125. [Google Scholar] [CrossRef] [Green Version]
  45. Szmigielski, R.; Surratt, J.D.; Gómez-González, Y.; Van der Veken, P.; Kourtchev, I.; Vermeylen, R.; Blockhuys, F.; Jaoui, M.; Kleindienst, T.E.; Lewandowski, M.; et al. 3-Methyl-1,2,3-butanetricarboxylic acid: An atmospheric tracer for terpene secondary organic aerosol. Geophys. Res. Lett. 2007, 34, L24811. [Google Scholar] [CrossRef]
  46. Spólnik, G.; Wach, P.; Rudziński, K.J.; Skotak, K.; Danikiewicz, W.; Szmigielski, R. Improved UHPLC-MS/MS methods for analysis of isoprene-derived organosulfates. Anal. Chem. 2018, 90, 3416. [Google Scholar] [CrossRef] [PubMed]
  47. Chen, Y.Z.; Zhang, Y.; Lambe, A.T.; Xu, R.S.; Lei, Z.Y.; Olson, N.E.; Zhang, Z.F.; Szalkowski, T.; Cui, T.Q.; Vizuete, W.; et al. Heterogeneous hydroxyl radical oxidation of isoprene-epoxydiol-derived methyltetrol sulfates: Plausible formation mechanisms of previously unexplained organosulfates in ambient fine aerosols. Environ. Sci. Technol. Lett. 2020, 7, 460. [Google Scholar] [CrossRef]
  48. Olson, C.N.; Galloway, M.M.; Yu, G.; Hedman, C.J.; Lockett, M.R.; Yoon, T.; Stone, E.A.; Smith, L.M.; Keutsch, F.N. Hydroxycarboxylic acid-derived organosulfates: Synthesis, stability, and quantification in ambient aerosol. Environ. Sci. Technol. 2011, 45, 6468. [Google Scholar] [CrossRef]
  49. Nestorowicz, K.; Jaoui, M.; Rudziński, K.J.; Lewandowski, M.; Kleindienst, T.E.; Spólnik, G.; Danikiewicz, W.; Szmigielski, R. Chemical composition of isoprene SOA under acidic and non-acidic conditions: Effect of relative humidity. Atmos. Chem. Phys. 2018, 18, 18101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Lockwood, A.L.; Shepson, P.B.; Fiddler, M.N.; Alaghmand, M. Isoprene Nitrates: Preparation, separation, identification, yields, and atmospheric chemistry. Atmos. Chem. Phys. 2010, 10, 6169. [Google Scholar] [CrossRef] [Green Version]
  51. Jaoui, M.; Corse, E.W.; Lewandowski, M.; Offenberg, J.H.; Kleindienst, T.E.; Edney, E.O. Formation of organic tracers for isoprene SOA under acidic conditions. Atmos. Environ. 2010, 44, 1798. [Google Scholar] [CrossRef]
  52. Schindelka, J.; Iinuma, Y.; Hoffmann, D.; Herrmann, H. Sulfate radical-initiated formation of isoprene-derived organosulfates in atmospheric aerosols. Faraday Discuss. 2013, 165. [Google Scholar] [CrossRef] [PubMed]
  53. Schone, L.; Schindelka, J.; Szeremeta, E.; Schaefer, T.; Hoffmann, D.; Rudziński, K.J.; Szmigielski, R.; Herrmann, H. Atmospheric aqueous phase radical chemistry of the isoprene oxidation products methacrolein, methyl vinyl ketone, methacrylic acid and acrylic acid—Kinetics and product studies. Phys. Chem. Chem. Phys. 2014, 16, 6272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Szmigielski, R. Evidence for C5 organosulfur secondary organic aerosol components from in-cloud processing of isoprene: Role of reactive SO4 and SO3 radicals. Atmos. Environ. 2016, 130, 14. [Google Scholar] [CrossRef]
  55. Wach, P.; Spólnik, G.; Rudziński, K.J.; Skotak, K.; Claeys, M.; Danikiewicz, W.; Szmigielski, R. Radical oxidation of methyl vinyl ketone and methacrolein in aqueous droplets: Characterization of organosulfates and atmospheric implications. Chemosphere 2019, 214, 1. [Google Scholar] [CrossRef] [PubMed]
  56. Shalamzari, M.S.; Kahnt, A.; Vermeylen, R.; Kleindienst, T.E.; Lewandowski, M.; Cuyckens, F.; Maenhaut, W.; Claeys, M. Characterization of polar organosulfates in secondary organic aerosol from the green leaf volatile 3-Z-hexenal. Environ. Sci. Technol. 2014, 48, 12671. [Google Scholar] [CrossRef]
  57. Shalamzari, M.S.; Vermeylen, R.; Blockhuys, F.; Kleindienst, T.E.; Lewandowski, M.; Szmigielski, R.; Rudziński, K.J.; Spólnik, G.; Danikiewicz, W.; Maenhaut, W.; et al. Characterization of polar organosulfates in secondary organic aerosol from the unsaturated aldehydes 2-E-pentenal, 2-E-hexenal, and 3-Z-hexenal. Atmos. Chem. Phys. 2016, 16, 7135. [Google Scholar] [CrossRef] [Green Version]
  58. Zhang, H.F.; Worton, D.R.; Lewandowski, M.; Ortega, J.; Rubitschun, C.L.; Park, J.H.; Kristensen, K.; Campuzano-Jost, P.; Day, D.A.; Jimenez, J.L.; et al. Organosulfates as tracers for secondary organic aerosol (SOA) formation from 2-methyl-3-buten-2-ol (MBO) in the atmosphere. Environ. Sci. Technol. 2012, 46, 9437. [Google Scholar] [CrossRef]
  59. Barbosa, T.S.; Riva, M.; Chen, Y.Z.; da Silva, C.M.; Ameida, J.C.S.; Zhang, Z.; Gold, A.; Arbilla, G.; Bauerfeldt, G.F.; Surratt, J.D. Chemical characterization of organosulfates from the hydroxyl radical-initiated oxidation and ozonolysis of cis-3-hexen-1-ol. Atmos. Environ. 2017, 162, 141. [Google Scholar] [CrossRef]
  60. Hettiyadura, A.P.S.; Jayarathne, T.; Baumann, K.; Goldstein, A.H.; de Gouw, J.A.; Koss, A.; Keutsch, F.N.; Skog, K.; Stone, E.A. Qualitative and quantitative analysis of atmospheric organosulfates in Centreville, Alabama. Atmos. Chem. Phys. 2017, 17, 1343. [Google Scholar] [CrossRef] [Green Version]
  61. Hettiyadura, A.P.S.; Al-Naiema, I.M.; Hughes, D.D.; Fang, T.; Stone, E.A. Organosulfates in Atlanta, Georgia: Anthropogenic influences on biogenic secondary organic aerosol formation. Atmos. Chem. Phys. 2019, 19, 3191. [Google Scholar] [CrossRef] [Green Version]
  62. Bondy, A.L.; Craig, R.L.; Zhang, Z.; Gold, A.; Surratt, J.D.; Ault, A.P. Isoprene-derived organosulfates: Vibrational mode analysis by Raman spectroscopy, acidity-dependent spectral modes, and observation in individual atmospheric particles. J. Phys. Chem. A 2018, 122, 303. [Google Scholar] [CrossRef] [PubMed]
  63. Glasius, M.; Bering, M.S.; Yee, L.D.; de Sá, S.S.; Isaacman-VanWertz, G.; Wernis, R.A.; Barbosa, H.M.J.; Alexander, M.L.; Palm, B.B.; Hu, W.; et al. Organosulfates in aerosols downwind of an urban region in Central Amazon. Environ. Sci. Process. Impacts 2018, 20, 1546. [Google Scholar] [CrossRef]
  64. Stone, E.A.; Yang, L.M.; Yu, L.Y.E.; Rupakheti, M. Characterization of organosulfates in atmospheric aerosols at four Asian locations. Atmos. Environ. 2012, 47, 323. [Google Scholar] [CrossRef]
  65. Worton, D.R.; Goldstein, A.H.; Farmer, D.K.; Docherty, K.S.; Jimenez, J.-L.; Gilman, J.B.; Kuster, W.C.; de Gouw, J.; Williams, B.J.; Kreisberg, N.M.; et al. Origins and composition of fine atmospheric carbonaceous aerosol in the Sierra Nevada Mountains, California. Atmos. Chem. Phys. Discuss. 2011, 11, 17071. [Google Scholar] [CrossRef] [Green Version]
  66. Kourtchev, I.; Fuller, S.; Aalto, J.; Ruuskanen, T.M.; McLeod, M.W.; Maenhaut, W.; Jones, R.; Kulmala, M.; Kalberer, M. Molecular composition of boreal forest aerosol from Hyytiälä, Finland, using ultrahigh resolution mass spectrometry. Environ. Sci. Technol. 2013, 47, 4069. [Google Scholar] [CrossRef]
  67. Kourtchev, I.; Godoi, R.H.M.; Connors, S.; Levine, J.G.; Archibald, A.T.; Godoi, A.F.L.; Paralovo, S.L.; Barbosa, C.G.G.; Souza, R.A.F.; Manzi, A.O.; et al. Molecular composition of organic aerosols in Central Amazonia: An ultra-high-resolution mass spectrometry study. Atmos. Chem. Phys. 2016, 16, 11899. [Google Scholar] [CrossRef] [Green Version]
  68. Jaoui, M.; Szmigielski, R.; Nestorowicz, K.; Kolodziejczyk, A.; Sarang, K.; Rudziński, K.J.; Konopka, A.; Bulska, E.; Lewandowski, M.; Kleindienst, T.E. Organic hydroxy acids as highly oxygenated molecular (HOM) tracers for aged isoprene aerosol. Environ. Sci. Technol. 2019, 53, 14516. [Google Scholar] [CrossRef]
  69. Kleindienst, T.E.; Jaoui, M.; Lewandowski, M.; Offenberg, J.H.; Lewis, C.W.; Bhave, P.V.; Edney, E.O. Estimates of the contributions of biogenic and anthropogenic hydrocarbons to secondary organic aerosol at a Southeastern US location. Atmos. Environ. 2007, 41, 8288. [Google Scholar] [CrossRef]
  70. Chan, M.N.; Surratt, J.D.; Claeys, M.; Edgerton, E.S.; Tanner, R.L.; Shaw, S.L.; Zheng, M.; Knipping, E.M.; Eddingsaas, N.C.; Wennberg, P.O.; et al. Characterization and quantification of isoprene-derived epoxydiols in ambient aerosol in the Southeastern United States. Environ. Sci. Technol. 2010, 44, 4590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Ding, X.; Wang, X.M.; Gao, B.; Fu, X.X.; He, Q.F.; Zhao, X.Y.; Yu, J.Z.; Zheng, M. Tracer-based estimation of secondary organic carbon in the Pearl River Delta, South China. J. Geophys. Res. Atmos. 2012, 117, D05313. [Google Scholar] [CrossRef]
  72. Guo, S.; Hu, M.; Guo, Q.F.; Zhang, X.; Zheng, M.; Zheng, J.; Chang, C.C.; Schauer, J.J.; Zhang, R.Y. Primary sources and secondary formation of organic aerosols in Beijing, China. Environ. Sci. Technol. 2012, 46, 9846. [Google Scholar] [CrossRef]
  73. Hu, D.; Bian, Q.; Li, T.W.Y.; Lau, A.K.H.; Yu, J.Z. Contributions of isoprene, monoterpenes, beta-caryophyllene, and toluene to secondary organic aerosols in Hong Kong during the summer of 2006. J. Geophys. Res. Atmos. 2008, 113, D22206. [Google Scholar] [CrossRef]
  74. Kleindienst, T.E.; Lewandowski, M.; Offenberg, J.H.; Edney, E.O.; Jaoui, M.; Zheng, M.; Ding, X.A.; Edgerton, E.S. Contribution of primary and secondary sources to organic aerosol and PM2.5 at SEARCH network sites. J. Air Waste Manag. 2010, 60, 1388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lewandowski, M.; Jaoui, M.; Offenberg, J.H.; Kleindienst, T.E.; Edney, E.O.; Sheesley, R.J.; Schauer, J.J. Primary and secondary contributions to ambient PM in the Midwestern United States. Environ. Sci. Technol. 2008, 42, 3303. [Google Scholar] [CrossRef]
  76. Kourtchev, I.; Copolovici, L.; Claeys, M.; Maenhaut, W. Characterization of atmospheric aerosols at a forested site in Central Europe. Environ. Sci. Technol. 2009, 43, 4665. [Google Scholar] [CrossRef] [PubMed]
  77. Lewandowski, M.; Piletic, I.R.; Kleindienst, T.E.; Offenberg, J.H.; Beaver, M.R.; Jaoui, M.; Docherty, K.S.; Edney, E.O. Secondary organic aerosol characterisation at field sites across the United States during the spring-summer period. Int. J. Environ. Anal. Chem. 2013, 93, 1084. [Google Scholar] [CrossRef]
  78. Fu, P.Q.; Kawamura, K.; Chen, J.; Li, J.; Sun, Y.L.; Liu, Y.; Tachibana, E.; Aggarwal, S.G.; Okuzawa, K.; Tanimoto, H.; et al. Diurnal variations of organic molecular tracers and stable carbon isotopic composition in atmospheric aerosols over Mt. Tai in the North China Plain: An influence of biomass burning. Atmos. Chem. Phys. 2012, 12, 8359. [Google Scholar] [CrossRef] [Green Version]
  79. Offenberg, J.H.; Lewandowski, M.; Jaoui, M.; Kleindienst, T.E. Contributions of biogenic and anthropogenic hydrocarbons to secondary organic aerosol during 2006 in Research Triangle Park, NC. Aerosol Air Qual. Res. 2011, 11, 99–108. [Google Scholar] [CrossRef]
  80. Wang, W.; Wu, M.H.; Li, L.; Zhang, T.; Liu, X.D.; Feng, J.L.; Li, H.J.; Wang, Y.J.; Sheng, G.Y.; Claeys, M.; et al. Polar organic tracers in PM2.5 aerosols from forests in Eastern China. Atmos. Chem. Phys. 2008, 8, 7507. [Google Scholar] [CrossRef] [Green Version]
  81. Hopke, P.K. Review of receptor modeling methods for source apportionment. J. Air Waste Manag. 2016, 66, 237. [Google Scholar] [CrossRef]
  82. Zhang, Y.; Sheesley, R.J.; Schauer, J.J.; Lewandowski, M.; Jaoui, M.; Offenberg, J.H.; Kleindienst, T.E.; Edney, E.O. Source apportionment of primary and secondary organic aerosols using positive matrix factorization (PMF) of molecular markers. Atmos. Environ. 2009, 43, 5567. [Google Scholar] [CrossRef]
  83. Wang, Q.Q.; He, X.; Huang, X.H.H.; Griffith, S.M.; Feng, Y.M.; Zhang, T.; Zhang, Q.Y.; Wu, D.; Yu, J.Z. Impact of secondary organic aerosol tracers on tracer-based source apportionment of organic carbon and PM2.5: A case study in the Pearl River Delta, China. ACS Earth Space Chem. 2017, 1, 562. [Google Scholar] [CrossRef]
  84. Li, R.; Wang, Q.Q.; He, X.; Zhu, S.H.; Zhang, K.; Duan, Y.S.; Fu, Q.Y.; Qiao, L.P.; Wang, Y.J.; Huang, L.; et al. Source apportionment of PM2.5 in Shanghai based on hourly organic molecular markers and other source tracers. Atmos. Chem. Phys. 2020, 20, 12047. [Google Scholar] [CrossRef]
  85. Lanzafame, G.M.; Srivastava, D.; Favez, O.; Bandowe, B.A.M.; Shahpoury, P.; Lammel, G.; Bonnaire, N.; Alleman, L.Y.; Couvidat, F.; Bessagnet, B.; et al. One-year measurements of secondary organic aerosol (SOA) markers in the Paris Region (France): Concentrations, gas/particle partitioning and SOA source apportionment. Sci. Total Environ. 2021, 757, 143921. [Google Scholar] [CrossRef]
  86. Hu, D.; Bian, Q.J.; Lau, A.K.H.; Yu, J.Z. Source apportioning of primary and secondary organic carbon in summer PM2.5 in Hong Kong using positive matrix factorization of secondary and primary organic tracer data. J. Geophys. Res. Atmos. 2010, 115, D16204. [Google Scholar] [CrossRef]
  87. Hettiyadura, A.P.; Xu, L.; Jayarathne, T.; Skog, K.; Guo, H.; Weber, R.J.; Nenes, A.; Keutsch, F.N.; Ng, N.L.; Stone, E.A. Source apportionment of organic carbon in Centreville, AL using organosulfates in organic tracer-based positive matrix factorization. Atmos. Environ. 2018, 186, 74. [Google Scholar] [CrossRef] [Green Version]
  88. Hu, W.W.; Campuzano-Jost, P.; Palm, B.B.; Day, D.A.; Ortega, A.M.; Hayes, P.L.; Krechmer, J.E.; Chen, Q.; Kuwata, M.; Liu, Y.J.; et al. Characterization of a real-time tracer for isoprene epoxydiols-derived secondary organic aerosol (IEPOX-SOA) from aerosol mass spectrometer measurements. Atmos. Chem. Phys. 2015, 15, 11807. [Google Scholar] [CrossRef]
  89. De Sa, S.S.; Palm, B.B.; Campuzano-Jost, P.; Day, D.A.; Newburn, M.K.; Hu, W.W.; Isaacman-VanWertz, G.; Yee, L.D.; Thalman, R.; Brito, J.; et al. Influence of urban pollution on the production of organic particulate matter from isoprene epoxydiols in Central Amazonia. Atmos. Chem. Phys. 2017, 17, 6611. [Google Scholar] [CrossRef] [Green Version]
  90. Freney, E.; Sellegri, K.; Chrit, M.; Adachi, K.; Brito, J.; Waked, A.; Borbon, A.; Colomb, A.; Dupuy, R.; Pichon, J.M.; et al. Aerosol composition and the contribution of SOA formation over Mediterranean forests. Atmos. Chem. Phys. 2018, 18, 7041. [Google Scholar] [CrossRef] [Green Version]
  91. Budisulistiorini, S.H.; Li, X.; Bairai, S.T.; Renfro, J.; Liu, Y.; Liu, Y.J.; McKinney, K.A.; Martin, S.T.; McNeill, V.F.; Pye, H.O.T.; et al. Examining the effects of anthropogenic emissions on isoprene-derived secondary organic aerosol formation during the 2013 Southern Oxidant and Aerosol Study (SOAS) at the Look Rock, Tennessee ground site. Atmos. Chem. Phys. 2015, 15, 8871. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Formation pathways of isoprene SOA at low-NOx. For simplicity, only dominant pathways are presented. Dimers and degradation products of methyltetrol organosulfates (e.g., C5-alkene triols) are not included (adapted from Surratt et al., 2010 [6]). The four stereoisomeric forms of the 2-methyltetrols are shown in the box; the 2-methyltetrols occur in two diastereoisomeric forms, 2-methylthreitol and 2-methylerythritol, and each of them has two enantiomers. Also shown are MW 212 and 214 organosulfates, formed by further processing of 2-methyltetrol-related MW 216 organosulfates.
Figure 1. Formation pathways of isoprene SOA at low-NOx. For simplicity, only dominant pathways are presented. Dimers and degradation products of methyltetrol organosulfates (e.g., C5-alkene triols) are not included (adapted from Surratt et al., 2010 [6]). The four stereoisomeric forms of the 2-methyltetrols are shown in the box; the 2-methyltetrols occur in two diastereoisomeric forms, 2-methylthreitol and 2-methylerythritol, and each of them has two enantiomers. Also shown are MW 212 and 214 organosulfates, formed by further processing of 2-methyltetrol-related MW 216 organosulfates.
Atmosphere 12 00728 g001
Figure 2. Formation pathway leading to 2-methyltetrols at high-NOx (adapted from Sato, 2008 [7], copyright (2008), with permission from Elsevier).
Figure 2. Formation pathway leading to 2-methyltetrols at high-NOx (adapted from Sato, 2008 [7], copyright (2008), with permission from Elsevier).
Atmosphere 12 00728 g002
Figure 3. Formation pathway of isoprene SOA via methacrolein at high-NOx. Abbreviations: MACR: methacrolein; MAE: methacrylic acid epoxide; 2-MG: 2-methylglyceric acid; MPAN: methacryloylperoxynitrate (adapted from Lin et al., 2013 [8]).
Figure 3. Formation pathway of isoprene SOA via methacrolein at high-NOx. Abbreviations: MACR: methacrolein; MAE: methacrylic acid epoxide; 2-MG: 2-methylglyceric acid; MPAN: methacryloylperoxynitrate (adapted from Lin et al., 2013 [8]).
Atmosphere 12 00728 g003
Figure 4. Total ion chromatogram illustrating the power of capillary GC/MS using prior trimethylsilylation for separating isoprene SOA markers obtained for a fine daytime HVDS filter sample collected during the transition period of the 2002 LBA-SMOCC campaign. (1, 2 and 3) C5-alkene triols; (4) malic acid; (5) 2-methylthreitol; (6) 2-methylerythritol; (7) mannosan; (8) levoglucosan; (9) 1,6-anhydro-β-D-glucofuranose; (10) arabitol; and (11) mannitol. Compound 4 (malic acid) is a marker for oxidation of unsaturated fatty acids; compounds 7, 8 and 9 (mannosan, levoglucosan, and 1,6-anhydro-β-D-glucofuranose) are markers for biomass burning, and compounds 10 and 11 (arabitol and mannitol) are markers for fungal spores (source: Claeys et al., 2010 [23]).
Figure 4. Total ion chromatogram illustrating the power of capillary GC/MS using prior trimethylsilylation for separating isoprene SOA markers obtained for a fine daytime HVDS filter sample collected during the transition period of the 2002 LBA-SMOCC campaign. (1, 2 and 3) C5-alkene triols; (4) malic acid; (5) 2-methylthreitol; (6) 2-methylerythritol; (7) mannosan; (8) levoglucosan; (9) 1,6-anhydro-β-D-glucofuranose; (10) arabitol; and (11) mannitol. Compound 4 (malic acid) is a marker for oxidation of unsaturated fatty acids; compounds 7, 8 and 9 (mannosan, levoglucosan, and 1,6-anhydro-β-D-glucofuranose) are markers for biomass burning, and compounds 10 and 11 (arabitol and mannitol) are markers for fungal spores (source: Claeys et al., 2010 [23]).
Atmosphere 12 00728 g004
Figure 5. GC/MS chromatograms illustrating the power of reversed-phase C18 HPLC and HILIC for separating MW 216 methyltetrol sulfates (i.e., extracted ion chromatograms for the m/z 215/97 ion pair (A) and 215/215 ion pair (B)) The chromatograms obtained in both modes were overlain on one graph and normalized to compare their relative intensities. The vertical dashed line separates the chromatograms that were obtained with different methods. Two isomers are detected in the C18 reversed-phase mode, while four of the six isomers are observed in the HILIC mode (reprinted with permission from Spólnik et al., 2018 [46]; copyright (2018) American Chemical Society).
Figure 5. GC/MS chromatograms illustrating the power of reversed-phase C18 HPLC and HILIC for separating MW 216 methyltetrol sulfates (i.e., extracted ion chromatograms for the m/z 215/97 ion pair (A) and 215/215 ion pair (B)) The chromatograms obtained in both modes were overlain on one graph and normalized to compare their relative intensities. The vertical dashed line separates the chromatograms that were obtained with different methods. Two isomers are detected in the C18 reversed-phase mode, while four of the six isomers are observed in the HILIC mode (reprinted with permission from Spólnik et al., 2018 [46]; copyright (2018) American Chemical Society).
Atmosphere 12 00728 g005
Table 1. Overview of molecular markers for isoprene SOA. Abbreviations: 2-MG, 2-methylglyceric acid; NOS, nitrooxy organosulfate; OS, organosulfate. The compounds are listed according to increasing MW.
Table 1. Overview of molecular markers for isoprene SOA. Abbreviations: 2-MG, 2-methylglyceric acid; NOS, nitrooxy organosulfate; OS, organosulfate. The compounds are listed according to increasing MW.
Molecular Structure
+ Chemical Name
(Trivial Name)
Chemical
Formula
MW
FigureAnalysis
Method
Selected
References
Atmosphere 12 00728 i001
3-methyl-3-butenetriol
+ tautomer (C5-alkene triol)
Atmosphere 12 00728 i002
3-methyl-2-butene-1,2,4-triol (C5-alkene triol)
C5H10O3
118
GC/MS[6,18,20]
Atmosphere 12 00728 i003
2,3-dihydroxy-2-methyl-
propanoic acid (2-MG)
C4H8O4
120
3GC/MS[8,17,21,22]
Atmosphere 12 00728 i004
2-methylbutane-1,2,3,4-tetrol
(2-methyltetrol)
C5HO4
136
1
2
GC/MS[1,6,10,17,19]
Atmosphere 12 00728 i005
sulfoxyethanoic acid
(glycolic acid OS)
C2H4O6S
156
LC/MS[39,40,47,48]
Atmosphere 12 00728 i006
2,3-dihydroxy-2-methyl-
butanedioic acid
(2-methyltartaric acid)
C5H8O6
164
GC/MS[1,68]
Atmosphere 12 00728 i007
2-sulfoxypropionic acid
(lactic acid OS)
C3H6O6S
170
LC/MS[39,40,47]
Atmosphere 12 00728 i008
3-oxo-2-sulfoxybutane-1-ol
C4H8O6S
184
LC/MS[40]
Atmosphere 12 00728 i009
2-hydroxy-2-methyl-
3-sulfoxypropanoic acid (2-MG OS)
C4H8O7S
200
3LC/MS[8,38,39,40]
Atmosphere 12 00728 i010
+ isomers
3-hydroxy-4-methyl-4-sulf-
oxy-2(3H)-dihydrofuranone
C5H8O7S
212
1LC/MS[41,42,46,47]
Atmosphere 12 00728 i011
+ isomers
2,3-dihydroxy-4-methyl-
4-sulfoxytetrahydrofurane
C5H10O7S
214
1LC/MS[41,46,47]
Atmosphere 12 00728 i012
3-methyl-2-sulfoxy-
butane-1,3,4-triol + isomers (2-methyltetrol OS)
C5H12O7S
216
1LC/MS[12,38,39]
Atmosphere 12 00728 i013
several isomers
(2-MG dimer)
C8H14O7
222
3GC/MS[21]
Atmosphere 12 00728 i014
2,3-dihydroxy-3-methyl-4-
sulfoxybutanoic acid
(2-methylthreonic acid OS)
C5H10O8S
230
LC/MS[49]
Atmosphere 12 00728 i015
2-hydroxy-2-methyl-
3-sulfoxybutanedioic acid
(2-methyltartaric acid OS)
C5H9O9S
244
LC/MS[49]
Atmosphere 12 00728 i016
3-methyl-4-nitrooxy-
3-sulfoxybutane-1,2-diol
– several isomers
(2-methyltetrol NOS)
C5H11O9NS
261
LC/MS[38,39]
Atmosphere 12 00728 i017
(2-MG dimer OS)
C8H14O10S
302
LC/MS[8,38]
Atmosphere 12 00728 i018
1,4-dinitrooxy-3-methyl-
3-sulfoxybutane-2-ol
- several isomers
(2-methyltetrol NOS)
C5H10O11N2S
306
LC/MS[38]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Claeys, M.; Maenhaut, W. Secondary Organic Aerosol Formation from Isoprene: Selected Research, Historic Account and State of the Art. Atmosphere 2021, 12, 728. https://doi.org/10.3390/atmos12060728

AMA Style

Claeys M, Maenhaut W. Secondary Organic Aerosol Formation from Isoprene: Selected Research, Historic Account and State of the Art. Atmosphere. 2021; 12(6):728. https://doi.org/10.3390/atmos12060728

Chicago/Turabian Style

Claeys, Magda, and Willy Maenhaut. 2021. "Secondary Organic Aerosol Formation from Isoprene: Selected Research, Historic Account and State of the Art" Atmosphere 12, no. 6: 728. https://doi.org/10.3390/atmos12060728

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