Next Article in Journal
Nitrous Oxide Emission Fluxes in Coffee Plantations during Fertilization: A Case Study in Costa Rica
Previous Article in Journal
Data Intelligence Model and Meta-Heuristic Algorithms-Based Pan Evaporation Modelling in Two Different Agro-Climatic Zones: A Case Study from Northern India
Previous Article in Special Issue
Secondary Organic Aerosol Formation from Isoprene: Selected Research, Historic Account and State of the Art
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Green Leaf Volatiles in the Atmosphere—Properties, Transformation, and Significance

by
Kumar Sarang
,
Krzysztof J. Rudziński
* and
Rafał Szmigielski
*
Institute of Physical Chemistry of the Polish Academy of Sciences, Kasprzaka Str. 44/52, 01-224 Warszawa, Poland
*
Authors to whom correspondence should be addressed.
Atmosphere 2021, 12(12), 1655; https://doi.org/10.3390/atmos12121655
Submission received: 21 October 2021 / Revised: 28 November 2021 / Accepted: 2 December 2021 / Published: 9 December 2021

Abstract

:
This review thoroughly covers the research on green leaf volatiles (GLV) in the context of atmospheric chemistry. It briefly takes on the GLV sources, in-plant synthesis, and emission inventory data. The discussion of properties includes GLV solubility in aqueous systems, Henry’s constants, partition coefficients, and UV spectra. The mechanisms of gas-phase reactions of GLV with OH, NO3, and Cl radicals, and O3 are explained and accompanied by a catalog of products identified experimentally. The rate constants of gas-phase reactions are collected in tables with brief descriptions of corresponding experiments. A similar presentation covers the aqueous-phase reactions of GLV. The review of multiphase and heterogeneous transformations of GLV covers the smog-chamber experiments, products identified therein, along with their yields and the yields of secondary organic aerosols (SOA) formed, if any. The components of ambient SOA linked to GLV are briefly presented. This review recognized GLV as atmospheric trace compounds that reside primarily in the gas phase but did not exclude their transformation in atmospheric waters. GLV have a proven potential to be a source of SOA with a global burden of 0.6 to 1 Tg yr−1 (estimated jointly for (Z)-hexen-1-ol, (Z)-3-hexenal, and 2-methyl-3-buten-2-ol), 0.03 Tg yr−1 from switch grass cultivation for biofuels, and 0.05 Tg yr−1 from grass mowing.

Contents

1. Introduction   2
2. Sources of GLV   6
 2.1. Reasons or Causes of Emissions   6
 2.2. Synthesis in Plants   6
 2.3. Emission Data   8
3. Physical Properties of GLV   15
 3.1. Solubility in Water and Partition Coefficients   15
 3.2. UV Spectra   21
4. Gas-Phase Chemistry   27
 4.1. Products and Mechanisms   27
  4.1.1. Reactions with OH   28
  4.1.2. Reactions with NO3   35
  4.1.3. Reactions with O3   38
  4.1.4. Reactions with Cl   44
  4.1.5. Gas-Phase Photolysis   51
 4.2. Gas-Phase Kinetics   52
  4.2.1. Gas-Phase Reactions with OH Radicals   53
  4.2.2 Gas-Phase Reactions with NO3 Radicals   58
  4.2.3. Gas-Phase Reactions with O3   61
  4.2.4. Gas-Phase Reactions with Cl Radicals   64
  4.2.5. Gas-Phase Photolysis   68
  4.2.6. SAR and LFPR Methods   69
5. Aqueous-Phase Kinetic   69
6. Multiphase and Heterogeneous Transformation   72
 6.1. Smog-Chamber Studies   72
 6.2. Ambient Aerosols   83
 6.3. Aqueous and Multiphase Mechanisms   83
7. Atmospheric Impact of GLV   89
 7.1. Atmospheric Lifetimes of GLV   89
 7.2. SOA Potential of GLV   93
8. Conclusions   94
Abbreviations   95
References   96

1. Introduction

Green leaf volatiles (GLV)—C5 and C6 aldehydes, alcohols, ketones, and esters—belong to inducible volatile organic compounds synthesized and emitted by green plants in response to mechanical wounding, herbivore attack, or abiotic stress [1,2,3,4,5,6,7,8]. They are synthesized rapidly via the lipoxygenase (LOX) pathway within minutes or seconds and can alert other plants fostering interplant communication [1,9]. Furthermore, they are responsible for the “green odor” of green leaves [10]. The first burst of volatiles often occurs when the leaves and stems are cut, followed by more intense and hours-long emissions from drying material [11]. Together with GLV, wounded plants emit methanol, acetaldehyde, acetone, butanone, and formaldehyde [12]. Plants also emit GLV upon the light-dark transition [13] and in high-light exposure [14].
The GLV include (Z)-3-hexenol, (Z)-3-hexenyl acetate, (E)-2-hexenal, (Z)-2-hexen-1-ol, (Z)-2-penten-1-ol and 1-penten-3-ol [12,13,15,16,17,18,19,20]. Jasmonic acid (JA), methyl jasmonate (MeJa), methyl salicylate (MeSa), and 2-methyl-3-butene-2-ol (MBO) also count as GLV [21,22]. However, they formally belong to other groups of plant volatiles [23,24]. Table 1 lists the compounds we included in this review as GLV (Table 1). Approximate yearly defense-initiated and cutting-drying emission of GLV in North America ranges from 1.5 to 2.6 Tg C yr−1 with estimated rates of 0.1 to 0.2 µg-C g−1 h−1 [25].
The global annual emission of all BVOC reached 1087 Tg yr−1 estimated using the Community Land Model (CLM4) integrated with the MEGAN2.1 framework for estimating fluxes of biogenic compounds between terrestrial ecosystems and the atmosphere [50]. Isoprene was 535 Tg yr−1, monoterpenes — 162 Tg yr−1, and other compounds—390 Tg yr−1. The modeling did not treat GLV as a separate group of compounds. It included some GLV in the stress VOC group (3-hexenal, 2-hexenal, 3-hexenol, 3-hexenyl acetate, hexanal, 1-hexenol, MeJa, and MeSa) and considered MBO individually. The emission of (Z)-3-hexenal was estimated as 4.9 Tg yr−1, of (Z)-3-hexenol as 2.9 Tg yr−1, and MBO as 2.2 Tg yr−1. The annual emission of MBO calculated with MEGAN and averaged throughout 1980–2010 was 1.6 ± 0.1 Tg yr−1 [51].
Although the global emission of GLV is relatively small, the significance of local and seasonal emissions of GLV due to such events as harvests or grass mowing can be high for local air quality and SOA burden.
The emission budget of GLV can increase further if the GLV treatments become introduced to the world of agriculture, horticulture, and forestry. Prospective solutions are being investigated and discussed. Treatment of tomato with (Z)-3-hexenol prepared it for better defense against Tomato yellow leaf curl virus transmitted by whitefly [52]. Spraying with (Z)-3-hexenyl acetate reduced the salt stress in peanuts by a complex genetic mechanism increasing the photosynthetic rate, plant height, and shoot biomass [53]. Post-harvest fumigation of strawberries with (E)-2-hexenal and (Z)-3-hexenal decreased the mold infection rate [54]. Maize seedlings treated with physiological amounts of (Z)-3-hexenyl acetate showed less damage from cold stress and increased growth compared to untreated seedlings [55]. (Z)-3-hexenyl butyrate successfully modified the pre-harvest ripening of grapes [56].
Once in the atmosphere, GLV immediately encounter several reactants like OH and NO3 radicals or O3, which are ready to interact with unsaturated carbonyl compounds and contribute to the formation of secondary organic aerosols (SOAs) [57,58].
The modeling, which is the only way to estimate the global production and burden of SOA, provides mixed results. The intercomparison of 31 global chemistry-transport and general circulation models showed that the global annual output of SOAs varied from 13 to 121 Tg yr−1, with a median value of 19 Tg yr−1 [59]. For the models which considered the semi-volatile character of SOAs, the global annual output of the SOAs varied from 16 to 121 Tg yr−1 with a much higher median of 51 Tg yr−1. The STOCHEM model predicted 46.4 Tg yr−1 of SOAs, formed globally from sesquiterpenes (29 Tg yr−1) and monoterpenes (162 Tg yr−1) [60]. The corresponding SOA burden was 0.45 Tg. Approximately 0.010 Tg (2%) originated from isoprene, 0.188 Tg (44%) from monoterpenes, 0.187 Tg (41%) from both monoterpenes and sesquiterpenes, and 0.041 Tg (10%) from sesquiterpenes alone. Those results are consistent with GEOS-CHEM modeling based on the volatility basis set, which predicted 36.2 Tg yr−1 and 0.88 Tg of SOAs, including 21.5 Tg yr−1 and 0.53 Tg of biogenic SOA; as well as 14.7 Tg yr−1 and 0.35 Tg of anthropogenic and biomass burning SOAs [61]. The SOA production increased when modeling covered: the semi- and intermediately volatile OC; new SOA yields from chamber experiments; wet and dry deposition of organic vapors; and the photolytic and heterogeneous loss of SOAs. The new production of 132.2 Tg yr−1 included 8.8 Tg yr−1 from anthropogenic and biomass burning emissions, 97.5 Tg yr−1 from biogenic emissions, and 25.9 Tg yr−1 from semi- and intermediately volatile OC emissions. However, the SOA burdens have not changed and included 0.08, 0.59, and 0.21 Tg, respectively. Gas-phase processing of isoprene alone was estimated to contribute 6–30 Tg yr−1 of SOAs [62] with an additional 2 Tg yr−1 from aqueous-phase processing [63]. The CESM model with CAM5.3 and CLM4 components predicted 55.7 Tg yr−1 of isoprene SOA (1.07 Tg burden) for climate and emission conditions of 2000 [64]. Against all the quantities mentioned above, a less formal prediction of 1–5 Tg C yr−1 of SOA produced from GLV [65] seems overestimated and needs re-evaluation.
Major reviews published in the last decade provide little information on a GLV role in atmospheric chemistry and SOA formation. A special issue of Trends in Plant Science (“Induced plant volatiles: from genes to climate change”) [66] discussed the BVOC emissions and their general relations with the environment with a focus on emissions [8,67,68]. The review on molecular identification of organic compounds in the atmosphere [69] mentioned the possibility of SOA formation from GLV based on four references. The extensive evaluation of the tropospheric aqueous-phase chemistry [70] included the rate constants for the aqueous-phase reactions with a few GLV, based on one work. The review on the atmospheric chemistry of oxygenated volatile compounds [71] provided the lifetimes of a few GLV due to the gas-phase reactions with OH, based on one reference. The opinion piece on plant-derived secondary organic material in the air [72] mentioned GLV as SOA precursors but did not go into any details.
The purpose of this review is to provide a practical background framework for designing future research on the role of GLV in atmospheric processes and specifically in the formation of SOAs.

2. Sources of GLV

2.1. Reasons or Causes of Emission

Plants release GLV when attacked by animals, insects, and microbes or exposed to abiotic stress or mechanical wounding [16,27,35,43,48,73,74]. For instance, trees and shrubs emitted (E)-2-hexenal upon wounding and in response to bacterial pathogenesis [6]. Emissions of Pinus sylvestris saplings infected with large pine aphid (Cinara pinea Mordviko) contained methyl salicylate and (Z)-3-hexenyl acetate [75]. Plants infected with pathogens probably emit methyl salicylate to activate disease resistance in neighboring plants and healthy tissues of the emitting plant [7]. Blande et al. [44] detected significant emission of MeSa from Betula pendula and Alnus glutinosa trees infested with Euceraphis betulae (Koch.) or Pterocallis alni (De Geer) aphids. The emission of volatiles depended quantitatively on the stress dose for both abiotic and biotic stressors [76].
Plants stimulated with GLV rapidly produced jasmonic acid (JA) and emitted anti-herbivore volatiles [34]. Plants pre-treated with GLV had more JA and sesquiterpenes when attacked by herbivores. The released GLV can inform the neighboring plant ecosystem of possible danger and prompt it to prepare its defense [4,5,6,34,77]. Experiments showed that intact undamaged corn seedlings produced JA and emitted sesquiterpenes when treated with GLV [34]. The GLV-primed defense initially suppresses the growth of plants, but after a few days, the growth significantly increases [78].
GLV can attract carnivores or parasitoids of herbivores that attack plants [79]. For instance, (Z)-3-hexenal attracted a chalcidoic wasp Encarsia formosa, a natural parasitoid of whitefly Bemisia tabaci, a herbivore pest of tomato plants [80]. Products of GLV reactions with ozone in the atmosphere also attract predators [81].
GLV emitted by plant seedlings can decrease their acceptability by herbivores (e.g., snails) [82]. Zhang et al. [41] showed that Ips typographus antennae responded strongly to green leaf alcohols (Z)-3-hexen-1-ol, 1-hexanol, and (E)-2-hexen-1-ol emitted by non-host plants such as Betula pendula, B. pubescens, and Populus tremula. Thus, the purpose of emission may be to direct the host-searching insect away.
(Z)-3-hexenyl butyrate controls the stomata closure to regulate CO2 and water vapor transfer under abiotic stress like drought [56].
(Z)-3-hexenol, (Z)-3-hexenal, and (Z)-3-hexenyl acetate were also emitted by velvet mosquito plants upon the light-dark transition, probably due to the dark activation of 13-lipoxygenase enzymes in chloroplasts [13].

2.2. Synthesis in Plants

The GLV listed in Table 1 differ in the pathways by which they are synthesized [23,83]. Compounds traditionally considered to be GLV—C6 aldehydes, alcohols, ketones, and esters, along with C5 compounds and MeJa—are produced via the LOX pathway. MBO is a hemiterpene synthesized by the mevalonate or DMAPP pathway [24]. MeSa is an aromatic compound synthesized by the shikimic acid pathway [23]. (Z)-3-hexenyl-propionate, (Z)-3-hexenyl butyrate, and (Z)-3-hexenyl isobutyrate are esters probably synthesized from (Z)-3-hexenol by AAT [30,74].
The LOX pathway converts free fatty acids (linolenic, hexadecatrienoic, and linoleic acid) into C6 GLV. Scheme 1 shows the reactions for linolenic acid [1,20,35,73,84]. This acid is oxidized to 13-hydroperoxodecatrienoic acid by molecular oxygen at position 13 in a reaction catalyzed by a non-heme iron-containing enzyme lipoxygenase (13-LOX). Then, hydroperoxyl lyase (HPL) converts the hydroperoxo acid into C6 aldehyde (Z)-3-hexanal and a C12 oxo acid (Z)-9-traumatin. The aldehyde is converted to (Z)-3-hexenol by alcohol dehydrogenase (ADH). The latter is converted to (Z)-3-hexenyl acetate by alcohol acyltransferase (AAT). In an alternative pathway, isomerase converts (Z)-3-hexanal to (E)-2-hexanal. The ADH enzyme converts the latter isomer to (E)-2 hexenol, which in turn is converted to (E)-2-hexenyl acetate by AAT. Hexadecatrienoic acid follows a similar pathway, including (Z)-7-dinortraumatin as a counterpart to oxo acid (not shown here).
Interestingly, hexenals can also form directly from membrane lipids, without their conversion to fatty acids by lipase [1,85]. Linoleic acid follows a path similar to Scheme 1 with hexanal and hexanol produced (not shown here).
Scheme 2 explains the formation of C5 GLV from 13-hydroperoxodecatrienoic acid [9]. First, lipoxygenase abstracts the OH group from OOH. The radical formed undergoes C5–C13 β-scission to give 1-penten-3-ol. The latter is converted to 1-penten-3-one by ADH.
13-hydroperoxodecatrienoic acid is also a precursor of JA and MeJa (Scheme 3) [9,32,49]. First, it is converted to alkene oxide by allene oxide synthase (AOS). Then, the oxide is cyclized to 12-oxophytodienoic acid by allene oxide cyclase (AOC). Finally, that acid is reduced and β-oxidized three times to give JA.

2.3. Emission Data

The variety and quantity of emitted GLV vary between plant species [84]. The emission rates range from values below the detection limits to 100 μg g−1 fresh weight. The fraction of some components increased with the total amount of emitted GLV, without any general patterns. Table 2 collects emission values and emission rates of GLV and their mixtures measured in the field and laboratory experiments. Emission measures were a GLV mass per mass of leaves or plant material, dry or fresh, or a GLV mass per land area. Quantities that refer to the leaf area are used in quantifying the VOC emission from living vegetation and modeling atmospheric chemistry and aerosol formation. Values referring to the mass of plant material help describe events like harvesting or grass mowing and compare VOC emission between species. Table 3 shows ambient concentrations of GLV and headspace concentrations measured in laboratory or field chambers.
Several researchers developed models that predict emission fluxes of volatile compounds, including GLV. The most comprehensive is the Model of Emission of Gases and Aerosols from Nature version 2.1 (MEGAN) [50,115], which aimed to estimate fluxes of biogenic compounds between terrestrial ecosystems and the atmosphere. MEGAN divided BVOC into 19 classes. There is no separate GLV class, but the “Stress VOC” class includes 3-hexenal, 2-hexenal, 3-hexenol, 3-hexenyl acetate, hexanal, 1-hexenol, MeJa, and MeSa. 2-Methyl-3-buten-2-ol is a separate class alone. MEGAN requires the input of meteorological, radiation, and land cover data, including the emission factors for 15 predefined classes of vegetation.
Heiden et al. [116] studied the emission dynamics of (Z)-3-hexenol, (Z)-3-hexenal, (E)-3-hexenol, (E)-2-hexenol, (E)-2-hexenal, hexanal, and 1-hexanol produced by various plants (tobacco, corn, pine, tomato, sunflower, broadleaf bean) after exposure to various stresses (ozone, pseudomonas infection, wounding) in a laboratory reactor. They obtained a general correlation:
ϕ ϕ m a x = A exp k 1 t exp k 2 t
where ϕ is the actual emission rate, ϕmax is the maximum emission rate in mol cm−2 s−1, A is the scaling factor to force ϕ = ϕmax at the correct experimental time, t is time in h, k1 and k2 measured in h−1 are empirical factors. For sunflower and broadleaf bean, the equation without the k1 term provided a good approximation. The parameters in Equation (1) are available in the original publication.
Kirstine and Galbally [19] developed a model for estimating BVOC emission from uncut and cut grass in urban environments. They estimated that hexenyl-type compounds (C6 GLV alcohols and aldehydes) constituted more than 70% of total BVOC emission upon initial wounding of grass and 22–40% during drying of cut material. For Sydney and Melbourne, Australia, the total emissions were 2.9 × 109 g yr−1 and 2.9 × 109 g yr−1, respectively, from uncut grass, 4.5 × 109 g yr−1 and 5.4 × 109 g yr−1 from cut grass, and 1.2 × 1011 g yr−1 and 2.0 × 1011 g yr−1 from other biogenic sources. The diurnal variation of BVOC flux varied from 0 in the night to peak values between 2 and 3 pm. Maximum emissions occurred in November and December.
Karl et al. [46] provided several model correlations for estimating MBO fluxes over vegetation, based on sensible heat (ωT) and latent heat fluxes (ωq), temperature (T), and two linear models (GLM1, GLM2) using temperature and light (PAR). The correlations utilized VOC fluxes measured at the Niwot Ridge AmeriFlux site in the Roosevelt National Forest in the Rocky Mountains of Colorado, USA, using the eddy covariance approach with PTR-MS instrument as a VOC sensor.
Geron et al. [47] provided a formula estimating the temperature dependence of MBO emission on temperature:
M = M S e β T T s
where M is the emission rate of MBO in mg C m−2 h−1 at temperature T in °C, Ms is the emission rate at standard temperature Ts = 30 °C, β is the temperature response factor (average β = 0.17, for the growing season β = 0.1).
Harley et al. [98] developed a model that described the diurnal emission of MBO and other BVOC from trees at varying temperatures and incident photosynthetic photon flux densities based on several fitted parameters like energies of activation and deactivation, and temperature and light scalars.

3. Physical Properties of GLV

3.1. Solubility in Water and Partition Coefficients

Henry’s constants, 1-octanol/water partition coefficients, and 1-octanol/air partition coefficients are collected in Table 4. Along with the experimental values, the table contains estimates obtained here using the EPI suite HENRYWIN v3.20 freely available from the US EPA [117]. EPI estimates Henry’s constants using three different methods (the bond estimation method, the group estimation method [118], and the VaporPressure/WaterSolubility estimation method. Estimates in Table 4 are the mean values from those three methods with standard deviations taken as uncertainties. The Log KOW procedure ver. 1.68 and Log Octanol-Air KOAWIN procedure ver. 1.1 were used to estimate the partition coefficients. The reviews of other estimation methods are provided elsewhere [119,120,121].
The octanol/water partition coefficient represents the equilibrium in a two-phase three-component system solute-water-octanol—it is a ratio of the solute concentration in the octanol-rich phase to the solute concentration in the water-rich phase [122]. It shows the balance between the lipophilicity and hydrophilicity of a solute. The octanol/air partition coefficient represents the equilibrium in a two-phase three-component system solute-air-octanol—it is a ratio of a solute concentration in octanol to the solute concentration in the air. It estimates the partitioning of a solute between air and various environmental matrices like soil, vegetation, and aerosol particles [135].
Table 5 presents the GLV solubility in water and equilibrium vapor pressure either determined experimentally or calculated using EPI MPBPVP and WSKOW suites [117].
The Henry’s constants in Table 4 attributed to Sander [120] are actually the mean values of the constants determined by several authors: for pentan-1-ol [118,119,138,143,144,145,146,147,148], hexanal [129,140,144,149,150], and nonanal [140,144,149,150]. One of the values for nonanal [149] seems an outlier compared to the other latest values and therefore was skipped in calculating the mean.
There are no experimental Henry’s constants for 1-penten-3-one (ethyl vinyl ketone). Still, the EPI estimate at 298 K in Table 4 is close to the mean empirical constant of (30.4 ± 10.1) M atm−1 determined for the structurally similar methyl vinyl ketone [120]. Similarly, the EPI estimate of the Henry’s constant for (Z)-2-pentenyl acetate at 298 K (Table 4) is close to the (2.7 ± 0.4) and (2.3 ± 0.4) M atm−1 values determined for the homolog amyl acetate and isoamyl acetate, respectively [120].
Hansel et al. [21] estimated Henry’s constant for several intermediate and final products of methyl jasmonate and methyl salicylate reactions with OH radicals (not shown here).
A few studies provided the temperature dependence of the GLV Henry’s constants, and most of them were for pentanol and hexanol. Gupta et al. [151] developed the classical van’t Hoff equations for dimensionless Henry’s constants based on empirical determination and valid between 313–363 K (Equations (3a) and (3b)):
l n K H ,   p e n t a n o l = 14.233 6559.6 T
l n K H ,   h e x a n o l = 11.705 5538.7 T
where KH was the ratio of the solute molar concentrations in the gas and aqueous phases.
Falabella et al. [152] correlated the experimental Henry’s constants determined in water and aqueous solutions of Na2SO4 for the temperature range 313–363 K using a modified theory of dilute solutions [153,154] (Equations (4a) and (4b)):
l n K H ,   p e n t a n o l = l n P H 2 O s a t + 10.04 T r 14.51 1 T r 0.355 T r + 2.72 e x p 1 T r T r 0.41 + 0.91 x
l n K H ,   h e x a n o l = l n P H 2 O s a t + 9.57 T r 10.61 1 T r 0.355 T r + 0.64 e x p 1 T r T r 0.41 + 1.11 x
where KH is in kPa, PsatH2O is the saturation pressure of water in kPa, Tr = T/Tcritical is the reduced temperature, and x is the concentration of Na2SO4 in mol kg−1 water.
Dohnal et al. [155] provided a correlation for the temperature dependence of the thermodynamic Henry’s constant for pentanol at 273–373 K (Equation (5)).
l n K H ,   p e n t e n o l o = 78.7049 99.5059 τ 97.8025   l n τ + D τ
where K0H is in kPa and τ = T/298.15.
Equation (6) approximates the temperature variation of the Henry’s Law constant for MBO in water [132]:
l n H MBO = 7230 ± 190 T 20.2 ± 0.7 ,   T = 275 295   K
where H is in M atm−1. The corresponding enthalpy of solvation is ΔHsol = −(60 ± 7) kJ mol−1. Equation (6) calculated the constants at 298 K and 303 K (58 and 38.9 M atm−1 resp.) that were lower than the experimental values from other authors (Table 4).
Zhou et al. [129] determined the temperature dependence of the Henry’s constant for hexanal (Equations (7a) and (7b)) and nonanal (Equations (8a) and (8b)) in freshwater and seawater across the range of 283–318 K:
log H hexanal, seawater = −8.35 + 2645/T
log H hexanal, freshwater = −8.76 + 2819/T
log H nonanal, seawater = −9.01 + 2555/T
log H nonanal, freshwater = −9.81 + 2929/T
where H is in M atm−1 and covers all possible hydrated forms of the solutes.
Liyana-Arrachchi et al. [156,157] simulated the behavior of MBO, MeSa, (Z)-3-hexen-1-ol, and (Z)-3-hexenylacetate at air-water interfaces using molecular dynamics approaches. The surface concentrations of GLV at the same bulk concentration decreased in the following order: (Z)-3-hexenyl acetate, (Z)-3-hexen-1-ol, MBO, and MeSa. The models accurately reproduced the experimental values of 1-octanol/water partition coefficients and the surface tension of solutions at 298 K and 0.01 MPa. The models indicated that all four GLV tended to remain at water-air interfaces with polar groups oriented towards the water. The observation indicated that the atmospheric reactions of those GLV may proceed at the aqueous interfaces rather than in the bulk gas or aqueous phases.

3.2. UV Spectra

UV spectroscopy is often used to follow the concentrations of reactants quantitatively in laboratory systems. The UV spectra of compounds quickly inform us about the possible photodegradation of these products. The energy flux of solar UV radiation is practically null below 290 nm [158], so compounds that do not absorb light above this value do not undergo photolysis in the atmosphere. The absorption cross-sections in the gas phase and the extinction coefficients in the aqueous phase were defined with Equations (9a) and (9b), resp. Formally, Equation (9c) converts both quantities between each other.
σ λ = l n I λ / I 0 λ L c
ε λ = l n I λ / I 0 λ L c = A λ L c
ε = σ N × 10 3
where σ (λ) is the absorption cross-section in cm2 molecule−1 at wavelength λ, ε(λ) is extinction coefficient in cm−1 M−1, L is the optical path in cm, C is the concentration of absorbing gas in molecule cm−3, c is the concentration of absorbing solute in M, I(λ) and I0(λ) are the light intensities in the presence and absence of absorbing gas, resp., A(λ) is the absorbance of a solute at λ, N is Avogadro’s number.
Jiménez et al. [57] determined the absolute UV absorption cross-sections of pure gaseous 1-penten-3-ol (1.5–6.5 mmHg), 1-penten-3-one (1.1–7.4 mmHg), and (Z)-3-hexen-1-ol (0.2–0.4 mmHg) in a Pyrex absorption cell with quartz windows at 298 K (Figure 1 and Figure 2). Jiménez et al. [159] used a similar setup to determine the UV absorption cross-sections of pure vapors of hexanal (0.9–7.0 mmHg) and (E)-2 hexenal (0.6–3.5 mmHg) at room temperature (Figure 3 and Figure 4).
Xing et al. [160] determined the UV absorption cross-section of gaseous (Z)-3-hexenal (0.2–7.0 mmHg) at 298 K using a Horiba Jobin Yvon, Triax-320 spectrophotometric setup (Figure 5). Absorption followed the Beer’s Law in the range 240–340 nm.
O’Connor et al. [161] determined the UV absorption cross-sections of gaseous (E)-2-hexenal and hexanal in a Pyrex cell with quartz windows at 297 ± 3 K (Figure 3 and Figure 4). They used both a static and a dynamic method. The cell was either filled with a given amount of aldehyde or flushed continuously with a mixture containing 1.7 × 1016–8.8 × 1016 molecule cm−3 of aldehyde and nitrogen. The cell pressure varied between 1.0 and 5.8 mmHg. The dynamic method failed for (E)-2-hexenal, probably because of leaks. The absorption cross-sections were slightly lower than [159], but the main absorption peaks appeared at the same wavelength.
Kalalian et al. [162] determined the UV absorption cross-sections for gaseous (E)-2-pentenal (0.2–9 mmHg) and (E)-2-hexenal (0.2–3 mmHg) at 298 K (Figure 1 and Figure 4). Their results agree well with [159,161] in terms of peak wavelengths and with [159] in terms of intensities. Absorption below 245 nm resulted from the π-π * transition and the absorption band between 270 and 370 nm–to the n-π* transition of the carbonyl group absorption.
Richards-Henderson et al. [22] determined the UV absorption coefficients for (Z)-3-hexen-1-ol, (Z)-3-hexenyl acetate, MeSa, MeJa, and 2-methyl-3-butene-2-ol in aqueous solutions of pH = 5.5 ± 03 at 298 K (Figure 2, Figure 6, and Figure 7). Their results agree well with [163] and this work regarding peak location, while differences in the intensities can result from experimental errors [163].
Canosa-Mas et al. [164] recorded a UV spectrum of MeSa vapors at equilibrium with liquid phase at 293 K (Figure 6d).
Sarang et al. [163] presented UV spectra of 1-penten-3-ol, (Z)-2-hexen-1-ol, and (E)-2-hexen-1-al in aqueous solutions at 298 K and pH = 7 (Figure 1, Figure 4 and Figure 8). The spectra were recorded with a double-beam Lambda 900 UV/VIS/NIR spectrometer (Perkin Elmer Instruments). We also included unpublished spectra of (E)-2-hexenal, (Z)-2-hexen-1-ol, (Z)-3-hexenyl acetate, MBO, and MeSa in aqueous solutions recorded at 295 K with a double-beam Jasco V-570 UV-VIS-NIR spectrophotometer using water as a reference medium (Figure 4, Figure 6, Figure 7, Figure 8).
Rudich et al. [165] determined the UV absorption cross-section of MBO in the gas phase using a flow-through cell at 10 mmHg to minimize the influence of isoprene formed by the decomposition of MBO (Figure 7a).
Figure 1. Absorption cross-sections σ of (a) 1-penten-3-one, (b) 1-penten-3-ol [57], and (c) (E)-2-pentenal [162] in the gas phase at 298 K, and absorption coefficients ε of (d) 1-penten-3-ol in the aqueous phase at 298 K and pH = 7 [163].
Figure 1. Absorption cross-sections σ of (a) 1-penten-3-one, (b) 1-penten-3-ol [57], and (c) (E)-2-pentenal [162] in the gas phase at 298 K, and absorption coefficients ε of (d) 1-penten-3-ol in the aqueous phase at 298 K and pH = 7 [163].
Atmosphere 12 01655 g001aAtmosphere 12 01655 g001b
Figure 2. (Z)-3-hexen-1-ol: absorption cross-sections σ in the gas phase [57] (a) and absorption coefficient ε in the aqueous phase at 298 K and pH = 5.5 [22] (b).
Figure 2. (Z)-3-hexen-1-ol: absorption cross-sections σ in the gas phase [57] (a) and absorption coefficient ε in the aqueous phase at 298 K and pH = 5.5 [22] (b).
Atmosphere 12 01655 g002aAtmosphere 12 01655 g002b
Figure 3. Absorption cross-sections σ of hexanal in the gas phase (a) [159], (b) at 297 K [161], and (c) at room temperature [166].
Figure 3. Absorption cross-sections σ of hexanal in the gas phase (a) [159], (b) at 297 K [161], and (c) at room temperature [166].
Atmosphere 12 01655 g003
Figure 4. (E)-2-hexenal: absorption cross-sections σ in the gas phase (a) at 298 K [159], (b) at 297 ± 3 K [161], ((c), top and middle plate) at 298 K [162], and (d) absorption coefficients ε in the aqueous phase at 298 K and pH = 5.5 Sarang [163] and this work.
Figure 4. (E)-2-hexenal: absorption cross-sections σ in the gas phase (a) at 298 K [159], (b) at 297 ± 3 K [161], ((c), top and middle plate) at 298 K [162], and (d) absorption coefficients ε in the aqueous phase at 298 K and pH = 5.5 Sarang [163] and this work.
Atmosphere 12 01655 g004aAtmosphere 12 01655 g004b
Figure 5. Absorption cross-sections σ of (Z)-3-hexenal in the gas phase at 298 K [160].
Figure 5. Absorption cross-sections σ of (Z)-3-hexenal in the gas phase at 298 K [160].
Atmosphere 12 01655 g005
Figure 6. Absorption coefficients ε of (a) (Z)-3-hexenyl acetate, (b) methyl jasmonate, and (c) methyl salicylate in aqueous solutions at 298 K and pH = 5.5 (Richards-Henderson [22]) and pH = 7 (this work), and (d) absorption cross-sections of methyl salicylate vapor at 293 K [164].
Figure 6. Absorption coefficients ε of (a) (Z)-3-hexenyl acetate, (b) methyl jasmonate, and (c) methyl salicylate in aqueous solutions at 298 K and pH = 5.5 (Richards-Henderson [22]) and pH = 7 (this work), and (d) absorption cross-sections of methyl salicylate vapor at 293 K [164].
Atmosphere 12 01655 g006aAtmosphere 12 01655 g006b
Figure 7. 2-methyl-3-buten-2-ol: absorption cross-sections σ of 2-methyl-3-buten-2-ol in the gas phase [165] (a) and absorption coefficients ε in the aqueous phase at 298 K and pH = 5.5 (Richards-Henderson [22]) and pH = 7 (this work) (b).
Figure 7. 2-methyl-3-buten-2-ol: absorption cross-sections σ of 2-methyl-3-buten-2-ol in the gas phase [165] (a) and absorption coefficients ε in the aqueous phase at 298 K and pH = 5.5 (Richards-Henderson [22]) and pH = 7 (this work) (b).
Atmosphere 12 01655 g007
Figure 8. (Z)-2-hexen-1-ol: absorption coefficients ε in the aqueous phase at 298 K and pH = 7 Sarang [163] and this work.
Figure 8. (Z)-2-hexen-1-ol: absorption coefficients ε in the aqueous phase at 298 K and pH = 7 Sarang [163] and this work.
Atmosphere 12 01655 g008

4. Gas-Phase Chemistry

GLVs undergo reactions with oxidants such as hydroxyl (OH) radicals, nitrate (NO3) radicals, ozone (O3), and Cl atoms in the atmospheric air.
Atkinson and Arey [167,168] reviewed the literature on mechanisms and kinetics of gas-phase atmospheric reactions of VOC and BVOC but included no data on GLV except hexanal and hexenols. Section 4 discusses detailed mechanisms and products of GLV reactions with OH, NO3, O3, and Cl. This review did not aim to cover the theoretical modeling of atmospheric reactions. Interested readers may see an excellent review by Vereecken et al. [169].

4.1. Products and Mechanisms

Volatile organic compounds react with radicals X following the general Scheme 4 [168,170], which begins with the formation of an alkyl radical by adding X to a double C=C bond or by hydrogen abstraction. The alkyl radical reacts with oxygen to form the alkylperoxy radical RO2. The fate of RO2 depends on available reactants. RO2 self-reacts with another RO2 and gives a carbonyl and alcohol or an alkoxy radical. The reaction with HO2 radicals leads to hydroperoxides, while the reaction with NO2− to the peroxynitrate. The reaction of RO2 with NO gives either the nitrate or the alkoxy radical, which reacts further to various products.

4.1.1. Reactions with OH

GLV reactions with OH radicals begin with the OH addition to the C=C bond or the hydrogen abstraction by OH. Almost all experimental works cited in this section illustrate their results using the addition mechanisms. However, Gai et al. [171] compared both channels in reactions of several hexenols with OH radicals at CCSD(T)/6-311++G(2d,2p)//BH&HLYP/6-31 G(d,p) level of theory. The hydrogen abstraction channel included 11 abstractions at all C sites and excluded the OH group. The addition channel generally appeared faster, and the branching ratios of the abstraction channel ranged from 0.16 to 0.59 (Table 6). The overall theoretical rate constants (abstraction + addition) differed by (−46) to 82% from the experimental rate constants determined by those authors (Section 4.2.1, Table 15). Table 7 and Table 8 show the products of GLV–OH reactions and the corresponding experiments.
On the contrary, Du and Zhang [172] performed quantum chemical modeling of the OH radical reaction with MBO. The reaction proceeded mainly through the equally fast additions of OH to two unsaturated C atoms in MBO and the formation of C4 and C3 adducts. Both intermediates reacted further with oxygen and finally gave two primary products: glycolaldehyde CH2OHCHO and propanone CH3COCH3. The minor products, formaldehyde HCHO and 2,3-dihydroxy-3-methyl propanal (CH3)2COHCHO, formed through the C4 adduct [172].
Among the experimentalists, only Davis et al. [173] considered the hydrogen abstraction channel in reactions of GLV aldehydes with OH (Scheme 5, Scheme 6, Scheme 7). First, an alkyl radical forms and reacts with molecular oxygen to give the peroxy radical (Scheme 5).
In the presence of NOx, the peroxy radical reacts with NO2 to produce a peroxynitrate or reacts with NO to afford an alkoxy radical by oxygen abstraction (Scheme 6). The alkoxy radical enters a series of reactions with O2 and NO that lead to the formation of an aldehyde and alkyl glyoxal (Scheme 7).
We explain the principles of the OH addition mechanism using (Z)-3-hexen-1-ol as an example. The references quoted may refer to other GLV but includes the same mechanism principle. First, the OH radicals add to the C=C bond producing two isomeric alkoxy radicals. Those radicals react with molecular oxygen to form two isomeric peroxy radicals (Scheme 8) [95].
In the absence of NOx, the peroxy radicals undergo a self-reaction that produces isomeric alkoxy radicals (Scheme 9), as shown for similar hexenals [174,175,176] and MBO [177].
If NO is present, it reacts with peroxy radicals either to subtract one oxygen atom and produce two alkoxy radicals [95,173,174,175,176,177,178,179,180,181,182] or to afford two organic dihydroxy nitrates by addition [178,180] (Scheme 10).
The alkoxy radicals can isomerize [180,183], decompose [95,174,175,176,180], or undergo hydrogen abstraction [175,176] (Scheme 11). The products formed include respectively: (i) 3,4-dihydroxyhexanal, (ii) propanal and 3-hydroxypropanal, and (iii) two dihydroxyhexanones and one trihydroxyhexanone.
Propanal can react further with OH, O2, and NOx to give peroxynitrates, formaldehyde, and CO2 [95].
MBO reacts with OH like (Z)-3-hexen-1-ol, and final products include glycolaldehyde, acetone, formaldehyde, hydroxymethylpropanal, and trihydroxybutanal. Scheme 12 shows the plausible reaction paths for MBO-derived alkoxy radicals [177,178,180,182].
Furthermore, the MBO-derived peroxy radicals can react with NO to give MW 165 dihydroxynitrates (Scheme 13) [180].
The OH radicals add to GLV aldehydes like they do to GLV alcohols and lead to aldehydes, alkylglycolaldehydes, glyoxal, and CO (Scheme 14) [173]. Yet, an alternative decomposition of the alkoxy radical leads to aldehyde, alkylglyoxal, and CO2 (Scheme 7).
A saturated GLV pentan-1-ol reacts with OH by the hydrogen abstraction from a CH2-group at position 2 and subsequent reaction with oxygen, leading to a β-hydroxyalkoxy radical. The latter can isomerize to 1,5-dihydroxy-2-pentanone, decompose to butanal, or get autoxidized to 1-hydroxy-2-pentanone [184]. Initial hydrogen abstraction at position 3 leads to 3-hydroxypropanal, and at position 4, to 5-hydroxy-2-pentanone and 4-hydroxypentanal.
Table 7. Products from gas-phase reactions of GLV with OH radicals.
Table 7. Products from gas-phase reactions of GLV with OH radicals.
GLVProductYield %Expt
Table 8
Ref.
Pentan-1-olPentanal40.5 ± 8.2X[184]
Butanal16.2 ± 3.7
Propanal8.1 ± 1.9
Etanal18.1 ± 4.2
Formaldehyde25.1 ± 1.3
5-hydroxy-2-pentanoneobserved
3-hydroxypropanalobserved
1-penten-3-olFormaldehyde35 ± 4I[185]
Glycolaldehyde47 ± 6
(Z)-2-penten-1-olFormaldehyde11 ± 2
Propanal91 ± 13
Glycolaldehyde87 ± 11
(E)-2-hexen-1-olButanalMainII[176]
(E)-3-hexen-1-olPropanalMainII[176]
37 ± 7III [175]
(Z)-3-hexen-1-olPropanalMainII[176]
58 ± 8IV [180]
74.6 ± 6.7VIII [183]
3-hydroxypropanal101 ± 24IV[180]
48 + 48/−24VIII[183]
3,4-dihydroxyhexanoic acidObservedIV[180]
Dihydroxynitrates MW 179
(see Scheme 10)
ObservedIV, VIII[180,183]
Hydroxycarbonyl MW 132
(see Scheme 11)
ObservedVIII[183]
(Z)-3-hepten-1-olButanal33 ± 3III[175]
(E)-2-hexenylacetateButanalObservedXI[186]
2-methyl-3-buten-2-olAcetone14.1 ± 0.2V[182]
52 ± 5 molarVI[179]
58 ± 4VII[178]
observedIV[180]
67 ± 5IX[177]
Formaldehyde9.3 ± 3.3V[182]
35 ± 4 molarVI[179]
29 ± 3VII[178]
observedIV[180]
33 ± 3IX[177]
Formic acid9.3 ± 3.3V[182]
Glycolaldehyde28.0 ± 2.8V[182]
50 ± 5 molarVI[179]
61 ± 9VII[178]
78 ± 20IX[177]
66 ± 2XIII a[187]
29 ± 5XIII a[187]
2-hydroxy-2-methylpropanal19 ± 7VII[178]
31 ± 4IV[180]
31 ± 11IX[177]
12 ± 2XIII a[187]
37 ± 7XIII b[187]
Dihydroxynitrates MW 165
(see Scheme 13)
observedIV[180]
Organic nitrates5 ± 2VII[178]
Glyoxal26XIII a[187]
37XIII b[187]
COobservedV[182]
CO2observedV[182]
Nonanal1-nitooxy octane40XII[188]
Table 8. Experiments used to determine the rate constants listed in Table 7.
Table 8. Experiments used to determine the rate constants listed in Table 7.
ExptPhotoreactorDetectionTemp.pOH SourceRef.
TypeVol.
dm3
Material Katm
IChamber47SteelLP FTIR2981Ethyl nitrite[185]
IIBag80TeflonGC-FID (SPME)2981H2O2[176]
IIIChamber1080QuartzLP FTIR2981H2O2[175]
IVChamber7000TeflonGC-FID, GC-MS, API-MS2960.97Methyl nitrite[180]
VChamber480TeflonLP FTIR2950.97H2O2[182]
VIChamber47SteelLP FTIR295 Methyl nitrite[179]
VIIChamber5800TeflonLP FTIR, GC-FID, GC-MS, API-MS/MS2981Ethyl nitrite[178]
VIIIChamber6500–7900TeflonGC-FID, GC-MS, API-MS2981Methyl nitrite[183]
IXLISA CRAC EUPHORE LP FTIR2981HONO,
n-propyl nitrite
[177]
XReactor480DuranLF FIR, GC-PI2980.97Methyl nitrite[184]
XIBag80TeflonGC-FID2981H2O2[186]
XIIChamber5000TeflonGC-FID2981Isopropyl nitrite[188]
XIII aChamber28,000TeflonLIP294–298 Cyclohexane [187]
XIII b Cyclohexane
NOx
Priya and Senthilkumar [189] theoretically studied the reaction of MeSa with OH radicals using DFT methods with B3LYP, M06-2X [50], and MPW1K [51] functionals, and 6-311++G(d,p) basis set. The dominating hydrogen abstraction channel was the abstraction of OH hydrogen (Scheme 15). The dominating addition channel was the OH addition at para position (Scheme 16).
On the contrary, Seif et al. [190] theoretically found that the addition at the ipso position had the largest branching ratio and the addition at the para position, the smallest one. The rate constants ki for the individual addition channels calculated at 298 K and 1 atm using RRKM theory at MN15-L/aug-cc-pVTZ level with Eckart tunneling were 4.65, 2.02, 1.01, 0.24, 3.27, and 0.57 108 M−1 s−1, respectively for the positions from 1 to 6. The hydrogen abstraction mechanism appeared negligible because of the high Gibbs potential.

4.1.2. Reactions with NO3

There are not many papers on the mechanisms of GLV reactions with NO3. General principles can emerge from the works on reactions with other alkenes [168,170]. The reaction begins with the NO3 addition to a double C=C bond which yields two isomeric alkyl radicals (Scheme 17). The following steps are like those in the GLV-OH mechanism. Fantechi et al. proposed a similar mechanism for the reaction of MBO with NO3 radicals (Scheme 18) [182]. That mechanism includes an additional decomposition of alkoxy radicals, which leads to acetone. More decomposition steps for MBO-derived alkoxy radicals were proposed (Scheme 19) [191].
Products obtained from reactions of GLV with NO3 are listed in Table 9, while Table 10 briefly describes the corresponding experiments.

4.1.3. Reactions with O3

The ozonolysis of alkenes generally begins with a primary ozonide formation by O3 addition to a double C=C bond, which decomposes to Criegee biradicals and carbonyl products (Scheme 20) [194,195,196,197]. Those products can recombine to form a secondary ozonide.
Criegee biradicals stabilize by a range of reactions that depend on the R substituents and are presented below for individual GLV.
The basic mechanism of the ozonolysis of (Z)-3-hexenol includes the formation of propanol, 3-hydroxypropanal, and Criegee intermediates that stabilize by hydration and dehydration, leading to propanoic and 3-hydroxypropanoic acids (Scheme 21) [92,198].
Scheme 22 shows additional hydroperoxide product channels. They begin with Criegee stabilization by isomerization and include OH-driven abstraction steps, which may be plausible [198]. Further schemes present more isomerization channels of hexenylacetates.
Scheme 23 shows the basic mechanism for O3 reaction with (Z)-3-hexenyl acetate [92,196,198,199,200]. The primary carbonyls formed are propanal and 3-oxypropyl acetate, while Criegee intermediates stabilize, leading to propanoic and 3-acetoxy propanoic acids.
Scheme 24 and Scheme 25 show the theoretically predicted stabilization pathways of Criegee intermediates, which afford several carbonyl products [196].
Scheme 26 shows the mechanism of the MBO ozonolysis. Smog chamber experiments confirmed some channels only at dry conditions [177,182].
Priya et al. [189] studied the ozonolysis of MeSa based on a simplified chemical mechanism. It contained the cycloaddition of ozone to the C5-C6 bond and abstraction of one oxygen by H2 (Scheme 27). They managed to estimate the rate constants for those steps that were close to the experimental values.
Products observed in GLV reactions with ozone are collected in Table 11. Table 12 briefly describes the corresponding experiments. The formation of SOA particles during ozonolysis of GLV and their composition are discussed in Section 6.1.

4.1.4. Reactions with Cl

The gas-phase reaction of Cl atoms with GLV follows a mechanism like that for OH radicals, which includes the Cl addition to a double bond or hydrogen abstraction by Cl at a C atom with the formation of HCl [179,204,205]. Quantum-mechanical modeling showed that the addition is the dominant pathway while the hydrogen abstraction cannot be excluded [205,206]. A few studies have focused on the GLV-Cl reaction products (Table 13 and Table 14).
Scheme 28 shows the Cl addition to 1-penten-3-ol at two possible positions accompanied by the formation of alkyl radicals that can either isomerize or react with O2 molecules to form the peroxy radicals [205].
The peroxy radicals formed can isomerize or enter a self-reaction to decompose (Scheme 29). The products of those reactions can further decompose or react with molecular oxygen to give stable products and release HO2 radicals (Scheme 30).
The hydrogen abstraction in 1-penten-3-ol occurs preferentially at C3 position, followed by the formation of peroxy radicals that self-react to form the alkoxy radicals. The latter can also self-react to afford 1-penten-3-one with the release of H2O2 (Scheme 31) [205].
Scheme 32 shows the reaction pathways for the alkoxy radicals derived from the Cl radical addition and hydrogen abstraction reactions of (Z)-2-penten-1-ol [205]. In both cases, acetaldehyde is a secondary product.
Shashikala and Janardanan [208] studied the reaction of (E)-2-hexenal with Cl radicals using the density functional theory. The geometry optimization used the MP2/6-31G * level of theory (B3), and later single point energies of all MP2/6-31G * optimized stationary points were computed using BHANDHLYP/6–311 + G** (B2) level of theory. They found that the Cl radical was more reactive towards (E)-2-hexen-1-al than other atmospheric oxidants (OH, O3, NO3), both in the Cl addition and H abstraction channels.
The reaction of MBO with Cl was studied experimentally [179,207] and theoretically at the MP2(full)/6-311G(d,p) and CCSD(T)/6-311þG(d,p) level of theory [206]. In the absence of NO, the Cl addition appeared more favorable than the hydrogen abstraction. The addition at the C4 position provided almost 50% of the products by the molar yield. The identification of glycolaldehyde and formic acid among the products revealed the formation of intermediate OH radicals [207]. Scheme 33 shows the plausible transformation of the alkoxy radicals produced from the initial reactions of MBO with Cl.
In NO’s presence, the peroxy radicals derived from the MBO reaction with Cl convert to alkoxy radicals (Scheme 34) [204]. Scheme 35 and Scheme 36 show further reactions of the alkoxy radicals.
Scheme 37 shows the reactions of the alkoxy radicals formed by the hydrogen abstraction from MBO [204].
Elizabeth Gaona-Colmán et al. [186] studied the Cl addition to (E)-2-hexenyl acetate. Scheme 38 shows the plausible transformation of the alkoxy radicals formed by that reaction.
Priya and Senthilkumar [209] theoretically studied the mechanism of Cl radical reactions with MeSa using density functional theory at B3LYP and M06-2X levels of theory with 6-311++G(d,p) basis set. They found that the reaction involved the hydrogen abstraction at the meta position throughout all courses. They concluded that the reaction of MeSa with a Cl radical via hydrogen abstraction from the aromatic ring followed by reaction with O2 results in the formation of methyl 3-peroxy-2-hydroxy benzoate radical intermediate via a barrierless reaction (Scheme 39).
The authors evaluated reactions of methyl 3-peroxy-2-hydroxy benzoate radicals with HO2, NO2, and NO. The reaction with NO was the most favorable, with the lowest energy barrier close to 6.7 kJ mol−1 (Scheme 40) [209].

4.1.5. Gas-Phase Photolysis

O’Connor et al. [161] analyzed the sunlight photolysis of (E)-2-hexenal, (Z)-3-hexenal and (E,E)-2,4-hexadienal in a Euphore smog chamber in dry purified air (0.05–1.0% RH and 286–294 K) using LP FTIR, GC-FID, and GC-MS with EI ionization. The samples were preconcentrated on glass beads at −160 °C, then rapidly desorbed onto the chromatographic column. The authors found that CO was the major product (34% yield) of the photolysis of (Z)-3-hexenal (which gave small amounts of products) according to the Norrish type 1 mechanism (Scheme 41). The (Z)-2-pentenyl radical form reacted with O2 to produce the pentenyl peroxy radical, which, in turn, self-reacted to a pentenyl oxy radical. The pentenyl oxy radical reacted with oxygen to finally give (Z)-3-pentenal, which showed in the experiments. On the other hand, (E)-2-hexen-1-al was reversibly converted to (Z)-2-hexen-1-al.
Tang and Zhu [166] studied the formation of HCO radicals from the gas-phase photolysis of n-hexanal (and n-heptanal) using cavity ring-down spectroscopy. Scheme 42 shows thermodynamically feasible dissociation pathways after UV excitation of n-hexanal. The radical yields ranged from 0.1 to 0.08 for 2–8 mmHg partial pressures of hexenal. Furthermore, they used mass spectrometry to identify several other products: acetaldehyde CH3CHO and butene C4H8 (yield 0.28–0.31); CO and pentane C5H12 (yield 0.20); C5H11, 2-methylcyclopentanol (yield 0.12); and 2-ethylcyclobutanol (yield 0.3).

4.2. Gas-Phase Kinetics

This section presents the rate constants for gas-phase reactions of GLV with OH and NO3 radicals, O3, and Cl atoms, and the photolysis rate constants. The rate constants were determined either by the relative or absolute approach. The relative method compares a GLV reaction with an oxidant to the oxidant reaction with another compound for which the rate constant is known. From that comparison, the method derives a relative rate constant for the GLV-oxidant reaction. The absolute method directly follows the decay of GLV and the oxidant to determine its absolute rate constant.

4.2.1. Gas-Phase Reactions with OH Radicals

Table 15 contains the rate constants for gas-phase reactions of GLV with OH radicals determined experimentally at single temperatures, Table 16 briefly describe the experiments carried out to determine the constants. Table 17 shows the Arrhenius parameters for reactions studied over temperature ranges, Table 18 briefly describe the experiments carried out to determine the constants. The OH radicals were generated using several methods: photolysis of H2O2 at 248, 252, or 254 nm; photolysis of H2O at 165 nm; photolysis of methyl nitrite CH3ONO at 300 nm (Equation (15)); photolysis of ethyl nitrite CH3CH2ONO; photolysis of HNO3 at 248 nm; photolysis of HONO at 355 nm; reaction of F atoms with H2O or H atoms with NO2.
Most of the reactions have negative activation energies, as their rate decreases with increasing temperature. The exception is methyl salicylate, which is the only aromatic compound studied. In some cases, evaluated theoretically, the additional channel of a GLV-OH reaction had a positive activation energy while the hydrogen abstraction channel had a negative one [189].
Gibilisco et al. [176] correlated the rate constants for reactions of OH radicals with unsaturated alcohols obtained experimentally ((E)-2-hexen-1-ol, (E)-3-hexen-1-ol, and (Z)-3-hexen-1-ol, Table 15 and from the literature with the energy of the highest occupied molecular orbital (HOMO) of those alcohols calculated with the Gaussian package:
ln kOH (cm3 molecule−1 s−1) = -(1.3 ± 0.1) EHOMO − (10.3 ± 1.3)
Based on their results and the literature data, Gibilisco et al. [175] correlated the rate constants for reactions of GLV with OH radicals against those for reactions with Cl atoms:
log kOH = (0.29 ± 0.04) log kCl − 10.8
Peirone et al. [217] showed that the SAR relation [229] based on group-reactivity factors provided an estimation close to the experimental rate constants (Table 15).
Basandorj et al. [226] showed that the gas-phase reactions of OH and OD radicals with 2-methyl-3-buten-2-ol were faster by 15% in the presence of O2. At temperatures higher than 335 K, that rate constant of those reactions depended on the total pressure and was analyzed for the dependence on the concentration of an inert gas component. The ab initio theoretical calculation supported the analysis. The second-order rate constant for OH radicals decreased with pressure and temperature according to Equation (12):
k I I = k 0 T M 1 + k 0 T M k T F c 1 1 + [ log ( k 0 T M / k T ) ] 2
where k0 is the thermolecular rate constant at the low-pressure limit, k is the rate constant at the high-pressure limit, Fc is the collisional broadening factor, and M is the inert gas. For k = 8.3 × 10−12 exp(610/T) cm3 molecule−1 s−1 at 100 mmHg [165] and Fc = 0.6, the temperature dependence of k0 was evaluated from the experimental data:
k 0 = 2.5   ± 7.4 × 10 32 exp 4250 ± 1150 T   cm 6 molecule 2 s 1
The kinetic Equation (13) was consistent with a Lindemann-Hinshelwood mechanism:
OH + MBO → HO − MBO*
OH + MBO → HO − MBO*
HO–MBO* + M → HO − MBO + M*
DFT calculations showed that the radical formed by the addition of OH to the internal carbon was more stable than that formed by the addition to the terminal carbon. Additionally, the authors provided geometries, energies, and vibrational frequencies of all structures considered.
Formation of OH radicals by photodissociation of methyl nitrite often used to generate OH radicals in laboratory experiments follows the mechanism below (e.g., [225]):
CH3ONO + hν → CH3O + NO
CH3O + O2 → CH2O + HO2
HO2 + NO → OH + NO2
Methyl nitrite can be synthesized by the dropwise addition of 50% sulfuric acid to a saturated solution of NaNO2 in methanol.

4.2.2. Gas-Phase Reactions with NO3 Radicals

Table 19 contains the rate constants for gas-phase reactions of GLV with OH radicals determined experimentally at single temperatures, Table 20 briefly describe the experiments in which the constants were determined. Table 21 shows the Arrhenius parameters for reactions studied over temperature ranges, Table 22 briefly describe the experiments in which the constants were determined. Methods used to generate the NO3 radicals include thermal decomposition of N2O5, the reaction of F atoms with HNO3, and the reaction of NO2 with O3.
Pfrang et al. [192,231] showed that relative rate constants for the gas-phase reactions of NO3 with (Z)-2-hexen-1-ol, (E)-3-hexen-1-ol, (Z)-2-penten-1-ol, and 1-penten-3-ol at 298 ± 3 K obtained using N2O5 as a source of NO3 were higher than the absolute rate constants obtained using off-axis continuous-wave cavity-enhanced absorption spectroscopy (CEAS) using the reaction of F atoms with HNO3 as a source of NO3. That was caused by slow reactions of N2O5 with GLV ((5.0 ± 2.8) × 10−19 cm3 molecule−1 s−1 for (Z)-2-penten-1-ol; (9.1 ± 5.8) × 10−19 cm3 molecule−1 s−1 for 1-penten-3-ol; and (3.1 ± 2.3) × 10−18 cm3 molecule−1 s−1 for (E)-3-hexen-1-ol).

4.2.3. Gas-Phase Reactions with O3

Table 23 contains the rate constants for gas-phase reactions of GLV with O3 determined experimentally at single temperatures, Table 24 describe the experiments carried to determine the constants. Table 25 shows the Arrhenius parameters for reactions studied over temperature ranges, Table 26 briefly describe the experiments carried to determine the constants. In each experiment, external devices generated ozone.
Gibilisco et al. [239] studied the gas-phase reactions of O3 with (Z)-3-hexen-1-ol, (E)-3-hexen-1-ol, and (E)-2-hexen-1-ol without any OH scavengers added because they assumed that OH radicals formed during ozonolysis affect both the major and the reference reactants and OH influence is self-canceling and marginal. That assumption is not exact because the amounts of reactants consumed by OH radicals add to the quantities consumed by O3 and thus are not self-canceling:
Δ R 1 ,   t o t a l Δ R 1 ,   O H Δ R 2 ,   t o t a l Δ R 2 ,   O H Δ R 1 ,   o z o n e Δ R 2 ,   o z o n e
Several authors showed that the rate constants calculated with DFT methods agreed well with the experimental constants [238,241]. Besides, Zhang et al. [241] found that the theoretical and empirical rate constants for gas-phase reactions of O3 with (Z)-3-hexenyl esters increased with the length of the ester group.

4.2.4. Gas-Phase Reactions with Cl Radicals

Table 27 contains the rate constants for gas-phase reactions of GLV with Cl radicals determined experimentally at single temperatures, Table 28 briefly describe the experiments carried out to determine the constants. Table 29 shows the Arrhenius parameters for reactions studied over temperature ranges, Table 30 briefly describe the experiments carried out to determine the constants.
Gibilisco et al. [249] correlated the rate constants for gas-phase reactions of Cl with various unsaturated VOC against the energy of the highest occupied molecular orbital (EHOMO) calculated with the Gaussian package (Equation (12)).
ln k   cm 3 molecule 1 s 1 = 0.3 ± 0.1 E H O M O 19.0 ± 0.4  

4.2.5. Gas-Phase Photolysis

Table 31 shows the experimental rates of GLV photolysis, absolute or relative to the NO2 photolysis rate. All constants were based on the measured UV spectra of GLV (Section 3.2). Table 32 provides more details.
Photolysis was an insignificant degradation path in the troposphere for 1-penten-3-ol and (Z)-3-hexen -1-ol [57], (Z)-3-hexen-1-al [161], MBO [165], and MeSa [227]. Photodegradation of (Z)-3-hexenal was faster than its reaction with O3 (1 × 1012 molecule cm−3) and slower than its reaction with OH (2 × 106 molecule cm−3). Substantial photodegradation occurred for (E)-2-pentenal and (E)-2-hexenal [162] as well as for 1-penten-3-one and hexanal [57].

4.2.6. SAR and LFPR Methods

We did not intend to review the methods of estimating the rate constants based on structure-activity or linear free energy relationships. The interested readers are referred to the original publications on the correlations for gas-phase reactions of organic compounds with OH radicals [222,229,253,254,255], with OH, O3, and NO3 radicals [215,255,256,257,258,259], with Cl radicals [215,222,245,260], and with O3 [255,261]. Besides, SAR correlations were proposed for some aqueous-phase reactions [262,263,264].

5. Aqueous-Phase Kinetics

There is little data published on the aqueous-phase reactions of GLV, so we present all findings in one section. Reaction mechanisms are discussed in Section 6.3, together with multiphase experiments. Rate constants determined are collected in Table 33, Table 34, Table 35.
Richards-Henderson et al. [22] studied the aqueous-phase oxidation of (Z)-3-hexen-1-ol, (Z)-3-hexenyl acetate, MeSa, MeJa, and MBO induced by OH radicals formed by the photodissociation of H2O2. Control experiments without H2O2 showed that the photodegradation of reactants did not occur. The total mass yield of SOA products was determined gravimetrically after blowing the post-reaction solutions to dryness with N2. Unreacted GLV in dry residues was determined chromatographically after the reconstitution of samples with water. Rate constants for reactions were determined using a competitive kinetics technique against sodium benzoate (Table 33).
Sarang et al. [163] determined the relative rate constants for the aqueous-phase reactions of 1-penten-3-ol, (Z)-2-hexen-1-ol, and (E)-2-hexenal with OH, NO3, and SO4 radicals (Table 33 and Table 34). They used the Laser Flash Photolysis-Laser Long Path Absorption (LFP-LLPA) technique over the 278–318 K temperature range. The rate constants weakly increased with temperature. The impact of diffusion of reactants on the rate constants was: 30–60% (reactions with OH radicals); 5–23% (reactions with SO4 radicals); and 0.2–8% (reactions with NO3 radicals).
Richards-Henderson et al. [265] determined the relative rate constants for reactions of several GLV with 3C * triplet state and 1O2 * singlet molecular oxygen generated by irradiation of organic chromophores—3,4-dimethoxybenzaldehyde (DMB) or methoxyacetophenone (MAP) and Rose Bengal, resp.—with simulated sunlight. They used a competitive kinetics method against the reference reactants phenol or syringol (for 3C *) and tyrosine or furfuryl alcohol (for 1O2 *). The results are in Table 33. The yields of SOA products were determined using the procedure applied in the experiments with OH radicals [22] (see above). All 1O2 * reactions and reactions of 3DMB * with hexenol and methylbutenol were too slow to produce measurable amounts of SOA.
Mael et al. [266] synthesized intermediate products of the oxidation of 1-metyl-3-buten-1-ol—(3,3-dimethyloxiran-2-yl)methanol (2-methyl-3-buten-1-ol 2,3-epoxide), 2-(oxiran-2-yl)propan-2-ol (2-methyl-3-buten-1-ol 3,4-epoxide), and 3-methylbutane-1,2,3-triol), and (2,3-dihydroxyiso-pentanol, DHIP)—to study their aqueous-phase reactions with sulfates and nitrates as well as hydrolysis thereof. Products obtained from reactions of both epoxides in 1 M solutions of sulfate included DHIP and its sulfuric acid esters (two isomers). Products of reactions of 3,4 epoxide in 1 M nitrate solutions included two isomers of nitric acid esters of DHIP (Section 6.3, Scheme 44).
In addition, those authors determined the rate constants for the following reactions: (i) acid (H+) catalyzed hydrolysis of 2,3-epoxide (4.075 ± 0.004) × 10−1 M−1 s−1; (ii) acid (H+) catalyzed hydrolysis of 3,4-epoxide (4.36 ± 0.13) × 10−3 M−1 s−1; (iii) neutral hydrolysis of tertiary sulfuric acid ester from 2,3-epoxide (4.42 ± 0.20) × 10−6 s−1; (iv) neutral hydrolysis of tertiary nitric acid ester from 3,4-epoxide (2.10 ± 0.14) × 10−3 s−1; (v) neutral hydrolysis of primary sulfuric acid ester from 3,4-epoxide < 2 × 10−7 s−1; (vi) neutral hydrolysis of primary nitric acid ester from 3,4-epoxide < 2 × 10−7 s−1; and D2SO4 esterification of DHIP (1.292 ± 0.047) × 10−4 s−1, probably at room temperature.
Noziere et al. [132] found that a chemical reaction accompanied the uptake of MBO in aqueous solutions of sulfuric acid (Section 3.1). The first-order rate constant for that reaction increased with the concentration of the acid:
n k = 0.19 ± 0.02 W 14.6 ± 1.3     s 1
where W = 40–62% is the weight concentration of sulfuric acid. The reaction product was assigned to the mass spectrometer signal m/z 87 recorded in solutions. The proposed chemical mechanism included a pinacol rearrangement of protonated MBO leading to 3-methyl butanone (Section 6.3, Scheme 50). Liu et al. [267] used a rotated wetted wall reactor coupled with TOF MS with single-photon ionization to study uptake and aqueous-phase reactions of 2-methyl-3-buten-2-ol with H2O2 (0.1–1.0% wt) in aqueous solutions of sulfuric acid (40–60% wt) at 293 K and 5–108 mm Hg total pressure. MBO was monitored as the MBO+-CH3 ion cluster (m/z 71). The uptake coefficient was defined as the probability of MBO loss (from the gas phase) per collision with the aqueous surface (Equation (19)).
γ = 4 kV ω A
where ω (m s−1) is the mean molecular speed of MBO, V (cm3) is the volume of the reaction zone, A (cm2) is the surface area of solution, and k (s−1) is the first-order rate constant for the removal of MBO from the gas phase defined by Equation (20).
l n S S 0 = kL v av
where S and S0 are MS signals of MBO with and without loss from the gas phase, L (cm) is the contact distance between the gas and the solution, and vav (cm s−1) is the average gas flow velocity. In the absence of H2O2, the uptake of MBO was reversible at 40% acid concentration, irreversible (reactive) at 60% acid, and mixed at 50% acid. The addition of H2O2 at concentrations higher than 1% substantially increased the reactive character of the uptake. The uptake coefficient increased with the concentration of both sulfuric acid and H2O2. Equations (21a)–(21d) approximated the original data.
l n γ = 0.1591 W 18.059 , R 2 = 0.9909 ,   a t   H 2 O 2 = 0
l n γ = 0.167 W 17.656 , R 2 = 0.9891 , a t   H 2 O 2 = 0.1   %   wt
l n γ = 0.2159 W 19.04 , R 2 = 0.9633 , a t   H 2 O 2 = 0.5   %   wt
l n γ = 0.2188 W 18.468 , R 2 = 0.9935 , a t   H 2 O 2 = 1.0   %   wt
where W (% wt) is the weight concentration of H2SO4.
The gas-phase products accompanying the MBO uptake were determined online using MS and GC approaches and offline using FTIR. Isoprene was the only gaseous product during MBO uptake in sulfuric acid solutions. The products identified during the uptake in the mixed solutions of H2O2 and H2SO4 were acetaldehyde, acetone, and a product observed online that decomposed into isoprene when analyzed offline. Scheme 44 (Section 6.3) includes the mechanism proposed by the authors.
Ren et al. [268] studied the aqueous-phase reactions of MBO with sulfate radical-anions generated from K2S2O8 either by the photodissociation at 254 nm or by thermal dissociation at room temperature. Due to essentially different rates of dissociation, the photo experiments lasted for 1 hour while the thermal experiments—up to 150 days. The same products were identified in both types of experiments, albeit in different proportions (Section 6.1, Tables 39 and 40 (organosulfates)). The reaction mechanism is discussed in Section 6.3.
Liyana-Arachchi et al. [269] analyzed the behavior of MBO molecules and OH radicals at an air-water interface using a molecular dynamics approach. They found that encounters of the species at the interface were sufficiently numerous to indicate a possibility of a chemical reaction.
Heath et al. [270] determined the pseudo-first-order rate constants for the reaction of MeJa with OH radicals in irradiated aqueous solution (bulk) and thin aqueous films. The OH radicals were generated by the photolysis of H2O2. The extent of the reaction was followed by measuring the concentration of MeJa with HPLC-UV/DAD. The rate of reaction in the films was more significant than in bulk and increased with decreasing film thickness (Table 36). The readers may doubt whether the illumination of the bulk solution and thin films was equally effective.
Hansel et al. [21,271] applied HPLC-ESI-MS techniques to identify several intermediate and final low-volatile products formed in the aqueous-phase oxidation of MeJa and MeSa by OH obtained by the photodissociation of H2O2 and proposed plausible formation mechanisms (Section 6.3, Schemes 47 and 48, resp.).
In summary, the rate constants for aqueous-phase reactions of GLV with OH radicals were high (~109 M−1 s−1) and close to the diffusional limit (Table 33). Rate constants for reactions with SO4 were a little lower but still close to the diffusional limit (Table 32). Reactions with NO3 were still slower, with rate constants of the order 108 M−1 s−1. Reactions with singlet molecular oxygen and excited triple-state carbon appeared the slowest, with rate constants ranging between 105 and 107 M−1 s−1. All reactions of GLV with radicals weakly accelerated with temperature had energy of activation between 4 and 20 kJ mol−1.

6. Multiphase and Heterogeneous Transformation

6.1. Smog-Chamber Studies

Products of GLV processing in smog chambers are listed in: Table 37 (photooxidation and ozonolysis of (Z)-3-hexen-1-ol), Table 38 (ozonolysis of (Z)-3-hexenyl acetate), Table 39 (photooxidation of 2-methyl-3-buten-2-ol and ambient samples), Table 40 (organosulfates from various experiments and ambient samples). One should remember that compounds identified in aerosol samples may have multiple precursors. The experiments are briefly described below.
Barbosa et al. [279] studied the formation of SOA from the hydroxyl radical (OH)-initiated oxidation and ozonolysis of (Z)-3-hexen-1-ol. They used either non-acidified or acidified sulfate seed aerosols under different relative humidity conditions (12–95% and 5–52%, resp.) and 291–300 K temperatures. The OH radicals were generated by the photolysis of isopropyl nitrite. SOA yields in OH–GLV experiments varied from 13 to 34.2 μg m−3 (0.8–2%) for acidic seeds and from 15.2 to 32.1 μg m−3 (0.9–2%) for nonacidic seeds, with weak maxima at intermediate RH (~30%). The dependence is biased by the variation of other parameters, notably the NO and NOx concentrations. SOA yields in O3–GLV experiments varied from 5 to 52 μg m−3 (0.5–5.1%) for acidic seeds and from 5 to 30 μg m−3 (0.5–2.9%) for nonacidic seeds, and were significantly higher at low RH (0.7~5% at RH = 4–5% vs. 0.5–1% at RH = 44–51%). They discovered the formation of several organosulfates (Table 40) and confirmed their presence in PM2.5 field samples (Section 6.2).
Hamilton et al. [65] studied the SOA formation during ozonolysis or OH initiated oxidation of (Z)-3-hexen-1-ol and (Z)-3-hexenylacetate in the European Photoreactor chamber in Valencia. They used isoprene as a reference compound. The photolysis of HONO generated the OH radicals. The authors used the LC-ITMS approach to identify the products contained in SOA particles collected on quartz-fiber filters [272]. Ozonolysis of (Z)-3-hexenyl acetate produced 3-acetoxypropanal, 3-acetoxypropanoic acid, 3-acetoxypropane peroxic acid, and a range of oligomers with ester and ether linkages (Table 38). Almost all oligomers contained a terminal acetate group, which prevented further oligomeric growth. Ozonolysis of (Z)-3-hexen-1-ol produced a vast number of oligomers originating from 3-hydroxypropanal and 3-hydroxypropanoic acid (Table 37), according to the mechanism in (Section 6.3, Scheme 51).
Aerosol mass yields in the OH–GLV experiments were 43.2 μg m−3 (3.1%) for (Z)-3-hexen-1-ol, 9.7 μg m−3 (0.93%) for (Z)-3-hexenylacetate, and 18.5 μg m−3 (1.2%) for isoprene (initial concentrations 449, 498, and 511 ppbv, resp.). Aerosol mass yields from ozonolysis of GLV were 854 μg m−3 (9.6%) for (Z)-3-hexen-1-ol and 800 μg m−3 (8.5%) for (Z)-3-hexenylacetate.
O’Dwyer et al. [201] observed SOA formation during ozonolysis of 1-penten-3-ol, (Z)-2-penten-1-ol and 1-penten-3-one at 298 K and atmospheric pressure in a 3910 dm3 smog chamber (see Section 4.1.3). They used dry conditions and no seed aerosol. A burst of new particles, 30–50 nm in diameter, was observed immediately after the ozone introduction (~180 s). The particles’ number decreased in time, but their size increased to 130–150 nm at ~1.2 hours. The aerosol mass yields were 0.136–0.166 for 1-penten-3-ol, 0.156–0.164 for (Z)-2-penten-1-ol, and 0.028–0.034 for 1-penten-3-one.
Hamilton et al. [272] studied the composition of SOA obtained from the ozonolysis of (Z)-3-hexenylacetate in the EUPHORE chamber at 292 K (average). FTIR followed the decay of hydrocarbons. The formation and growth of SOA was followed by a scanning mobility particle sizer, condensation particle counter, and differential mobility analyzer. Samples of SOA were collected on quartz fiber filters using a high-volume pump. The filters were extracted by sonication with water, methanol, tetrahydrofuran, or acetonitrile. The extracts were evaporated to dryness and reconstituted with a water/methanol solvent (50:50 vol) for LC-MS/MS analysis with Li+ cationization. The major products identified are listed in Table 38. Fifteen other oligomers were distinguished, all with 3-acetoxypropanoic acid monomer units.
Jain et al. [198] observed the formation of SOA from ozonolysis of (Z)-3-hexen-1-ol, (Z)-3-hexenyl acetate, or effluents from grass clippings (Festuca, Lolium, and Poa) in a 775 l Teflon chamber at 298 ± K. Ozone was generated externally with a corona discharge apparatus. The initial molar ratio of O3 and GLV was 1:1. They measured the particle number and mass size distribution using a scanning mobility particle sizer. In grass clipping experiments, SOA formed immediately after ozone injection into the chamber. The mass loading of SOA reached 16 μg m−3 from (Z)-3-hexenyl acetate and 12 μg m−3 from (Z)-3-hexen-1-ol. The addition of propionaldehyde enhanced the SOA formation indicating it was a precursor of condensing products (Section 6.3, Scheme 51). They analyzed the particle composition using near-infrared laser desorption/ionization mass spectrometry. Compounds observed in SOA from grass clipping experiments included first-generation products (m/z 70–150) and later products (m/z 150–300).
Fischer et al. [48] determined the (1.03 ± 0.07%) SOA mass yields from gas-phase ozonolysis of 1-octen-3-ol in two Teflon chambers (see Section 4.2.3) equipped with scanning mobility particle sizers and electrical low power impactors. Ozone and 1-octen-3-ol mixing ratios were 200 ppb. They also observed SOA formation from ozonolysis of volatiles emitted in situ from cut sugarcane. The sugarcane volatiles included toluene, 3-hexenal, 2-hexenal, (Z)-3-hexenyl acetate, (Z)-3-hexen-1-ol, and 1-octen-3-ol. The gas-phase ozonolysis products included methyl ester of butanoic acid, heptanal, octanal, nonanal, and decanal, all identified against the NIST spectral library. The maximum SOA concentration observed was 1.6 ± 0.2 μg m−3. Several particle-phase products were identified using NIR-LDI-AMS (Section 6.3, Scheme 53). The SOA particles had a non-liquid character shown by the bounce-factor determinations.
Faiola et al. [295] investigated the SOA formation during dark ozonolysis and OH-induced photooxidation of Pinus sylvestris emissions. They compared the emissions from healthy plants and plants infested by large pine aphids (Cinara pinea Mord.). The initial concentration of MeSa was one of the parameters measured. SOA yield from ozonolysis ranged from 6.3 to 14.6% and did not depend on the infestation. SOA yield from the OH photooxidation of the emissions was: 10.5–23.2% for healthy plants and 17.8–26.8% for the aphid-infested plants.
Waza [296] studied the ozonolysis of (E)-2-hexenal, (Z)-3-hexen-1-ol, and (Z)-3-hexenylacetate in a rectangular 9.8 m3 Teflon-film chamber at ambient temperature and pressure and 50% RH. He observed the formation and growth of SOA particles. The SOA yields were smaller than those observed for α-pinene in comparable experiments.
Joutsensaari et al. [297] demonstrated the formation and growth of SOA particles from ozonolysis of emissions of white cabbage cultivars sprayed with MeSa in an illuminated 2.6 m3 growth chamber. The emissions contained terpenes, sesquiterpenes, (Z)-3-hexen-1-ol, (E)-2-hexenal, and (Z)-3-hexenyl acetate.
Shalamzari et al. [294] studied the growth of SOA on seed aerosol during the ozonolysis of (Z)-3-hexenal in a 14.5 m3 steel chamber with Teflon-coated walls. Relative humidity and acidity of seed aerosols varied. The OH radicals were generated from the NOx-mediated photooxidation of organic precursors. SOA samples were collected on Teflon-impregnated glass fiber filters. The methanol extracts of the filters were obtained using sonication, preconcentrated with a rotary evaporator, filtered, and evaporated to dryness with nitrogen stream. The residue was reconstituted with methanol, dried again, reconstituted with methanol/water solvent (1:1 col), and analyzed with UPLC-MS using reversed-phases or ion-pairing techniques. A significant MW 226 organosulfate was identified as a SOA-bound component from the ozonolysis of (Z)-3-hexenal (Table 40).
In a similar study, Shalamzari et al. [289] studied SOA from ozonolysis of (E)-2-pentenal, (Z)-3-hexenal, and (Z)-2-hexenal. They identified the following organosulfates, which could be related to (E)-2-pentenal: 3-sulfooxy-2,4-dihydroxy pentanoic acid (MW 230); 2-sulfooxy-3-hydroxy pentanoic acid and 3-sulfooxy-2-hydroxy pentanoic acid (MW214); lactic acid sulfate (MW 170); and 1-sulfooxy-2-hydroxy butane (MW 170) (Table 40). Those organosulfates were also produced during ozonolysis of (Z)-3-hexenal and (Z)-2-hexenal. 3-sulfoxy-2-hydroxy pentanoic acid was also formed during aqueous-phase sulfation of (E)-2-pentenoic acid with sulfate radical anions. The formation mechanisms of those products were suggested.
Harvey et al. [92] observed the formation of SOA from ozonolysis of (Z)-3-hexenol, (Z)-3-hexenylacetate, their mixtures, and emissions from turf grass (Festuca, Lolium, and Poa). They used a 775 l Teflon chamber at 296 K. Using a scanning mobility particle sizer, they measured total aerosol mass, aerosol particle number, and mass size distribution. The recorded SOA yields were 0.26 ± 0.01 and 0.24 ± 0.08%, respectively. For each GLV, the products identified included: propanoic acid, propenoic acid, acetic acid, propanal, and 2-propenal (from each GLV). Acetaldehyde was produced only from hexenyl acetate.
Carrasco et al. [177] studied the SOA formation during the gas-phase reactions of MBO with O3 and with OH radicals in the presence of NO. Experiments with ozone were carried out in the CRAC chamber, while experiments with OH–in the 204 m3 EUPHORE outdoor chamber with natural sunlight illumination (Section 4.2, Table 16 and Table 24). The chambers had multireflection optical systems coupled to FTIR spectrometers for monitoring the concentrations of gaseous reactants and products and the analyzers for monitoring O3, NOx, and NOy. Scanning mobility analyzers monitored the number, concentration, and size distribution of SOA particles. No significant yields of SOA occured in the reactions with OH radicals. Ozonolysis produced 9–25 μg m−3 aerosol mass at wet conditions (RH = 20–30%) and 74 μg m−3 at dry conditions equivalent to 0.3–1.5% and 1.8% yields, resp.
Chan et al. [187] studied SOA formation during photooxidation of 2-methyl-3-buten-1-ol initiated by OH radicals at 293–298 K, 4–66% RH, and high NOx or low NOx conditions in 28 m3 Caltech dual indoor chambers. The OH radicals were generated by the HONO photooxidation at high NOx and H2O2 photooxidation at low NOx. GC-FID, chemical ionization MS, and laser-induced phosphorescence were used to determine the concentration of 2-methyl-3-buten-1-ol, gas phase products, and glyoxal, resp. The mass yield of SOA was 0.001–0.0014 in the low NOx experiments and lower than 0.001 in the high NOx experiments.
Jaoui et al. [273] used a 14.5 m3 steel chamber with Teflon coated walls to study the photooxidation of 2-methyl-3-buten-2-ol initiated by OH radicals in the presence and absence of NOx and SO2. They generated OH radicals by the in situ photolysis of H2O2 or methyl nitrite and added ammonium sulfate seed aerosol to promote the condensation of low volatile products. In some experiments, they added SO2 as well. The reactants and gas-phase carbonyl products were determined using GC-FID and GC-MS. The aerosol particles were collected on quartz filters for further analysis of composition and EC/OC contents. The samples were extracted from filters, derivatized with BSTFA, and analyzed using GC-MS approaches. Size distribution, volume, and total number density of aerosol particles were measured during experiments with a scanning mobility particle sizer. Besides, samples of ambient PM2.5 aerosol were collected on quartz filters in the Research Triangle Park, NC USA, extracted, derivatized, and analyzed for composition. Conversion of MBO in all experiments was 70%, the yield of SOA was 0.7% (one experiment), and the yield of secondary organic carbon (SOC) ranged from 0.2 to 0.9%. The gas products identified included formaldehyde, acetaldehyde, glyoxal, glycolaldehyde, methylglyoxal, glyoxylic acid, 2,3butanedione, 2-oxopropane-1-3-dial, 2-hydroxy-2-methylpropanal, 2-hydroxypropane-1,3-dial, 2,3-dioxobutanal, 2,3-dioxobutane-1,4-dial, 3-hydroxy-2-oxo-isopentanol, 2,3-dihydroxy-3-methylbutanal, 2,3-dihydroxy-2-methylbutane dialdehyde, and 2-oxovaleric acid. Table 39 shows products identified in the particle phase.
Zhang et al. 2012 [293] studied the photooxidation of 2-methyl-3-buten-2-ol initiated by OH radicals in the presence of low NO in a 274 m3 outdoor smog chamber with natural sunlight. The OH radicals were generated from H2O2. Neutral or acidic seed aerosol was prepared from ammonium sulfate and sulfuric acid. Aerosol was sampled on glass-fiber filters. The samples were extracted with methanol and analyzed using LC-MS techniques. The SOA yield increased with the acidity of seed aerosol up to 1%. The main product identified–MW 200 organosulfate–was also determined in ambient SOA (Section 6.2). The same product was identified in aerosol samples from smog-chamber experiments with the uptake of (3,3-dimethyloxiran-2-yl) methanol (a 2-metyl-3-buten-1-ol epoxide) on seed aerosols carried out to confirm the epoxide pathways of SOA formation [298].
Novelli et al. [274] studied the photooxidation of MBO by O3 in the presence of NO in the 270 m3 outdoor chamber SAPHIR. They followed the evolution of OH and HO2 radicals and showed it was consistent with box modeling based on the MCM 3.3.1 mechanism. The hydrogen-shift pathways suggested by [299] were insignificant due to relatively low reaction rates. The major products observed in the chamber were acetone and formaldehyde.

6.2. Ambient Aerosols

Many compounds linked to the atmospheric transformation of GLV were firmly identified in ambient aerosol samples. In this section, we present the results from campaigns that intentionally regarded the reactions involving GLV.
Barbosa et al. [279], in a study supporting their chamber experiments, discovered several organosulfates in PM2.5 samples collected on prebaked quartz filters in the Atlantic Forest area (open tropical rainforest type) of the Botanical Garden in Rio de Janeiro in 2016 (Table 40). They identified organosulfates also in the companion smog-chamber experiments.
Ambient aerosol samples collected in K-puszta in summer 2003 contained 2,3-dihydroxypentanoic acid, which could originate from 2-pentenal [300], a product of (Z)-3-hexenal [161] photolysis.
Shalamzari et al. [294] analyzed PM2.5 samples collected in K-puszta from May to June 2006 (BIOSOL campaign). They used high-volume dichotomous PM2.5 samplers with quartz-fiber filters and sampled separately during days and nights. Punches from the filters were extracted with methanol, preconcentrated, filtered, and evaporated to dryness. The residues were reconstituted in methanol, divided in portions, and evaporated to dryness. Before the UPLC-MS analysis, the residues were reconstituted in methanol/water (1:4 vol) solvent. A characteristic MW 226 organosulfate H2C=C-CH(OH)-CH(OSO3H)-CH2-C(O)OH was identified in the samples, which could form by ozonolysis of (Z)-3-hexenal as shown in companion chamber experiments (Section 6.1, Table 40).
Shalamzari et al. [289] continued the analysis of K-puszta aerosol samples. They identified the following organosulfates linked to GLV in the companion chamber studies (Section 6.1, Table 40): 3-sulfoxy-2,4-dihydroxypentanoic acid (MW 230), 2-sulfoxy-3-hydroxypentanoic acid (MW 214), lactic acid sulfate (MW 170), 1-sulfooxy-2-hydroxybutane (MW 170).
Zhang et al. [293] collected ambient PM2.5 aerosol samples using high-volume samplers in Manitou Forest Observatory in the Rocky Mountains in 2011. The samples were extracted from filters with methanol and analyzed using LC-MS techniques. The MW 200 organosulfate was identified, which also formed from MBO in companion smog-chamber experiments (Section 6.1, Table 40). The compound was suggested as a plausible marker of MBO SOA.

6.3. Aqueous and Multiphase Mechanisms

GLV and products of their gas-phase reactions (Section 4.1) can partition into aqueous solutions or particle aqueous-phases. They can hydrolyze or react with other dissolved species by various mechanisms reviewed in this section.
GLV epoxides form by the oxidation of GLV in the gas phase. Scheme 43 shows the formation of MBO epoxides, including the 1,5-hydrogen shift and elimination of HO2 radical [266,298].
In the aqueous phase, GLV epoxides can hydrolyze or react with acids or acidic anions like sulfate or nitrate as reported for MBO (Scheme 44) [266,267], (Z)-3-hexenal [294], and (E)-2-pentenal [289]. Hydrolysis is acid-catalyzed [267].
GLV-derived hydroperoxides form during the ozonolysis or photooxidation of GLV in the gas phase [279]. After partitioning to the aqueous phase, they can react with acids (Scheme 45).
Aqueous-phase reactions of GLV with radicals and radical-anions can proceed by the addition of a radical to a C=C double bond, electron transfer, or hydrogen abstraction. The addition mechanisms lead to many products and are most interesting. The mechanism involving sulfate radical-anions and oxygen was proposed for isoprene [301,302], methyl vinyl ketone [303], (Z)-2-pentenoic acid [289], and recently for olefins, including MBO [268]. Scheme 46 shows the reaction mechanism for MBO based on the latter citation. The reaction starts with the addition of SO4 to the double C=C bond in MBO and the formation of carbon-centered alkyl sulfate radical (two isomers). Oxygen molecule adds to a radical center to form the peroxyalkyl sulfate radical, which can enter a self-reaction [181] and turn into the sulfate butanediol and hydroxybutanone sulfate products. In an alternative path, the peroxyradical loses one oxygen atom to some available reducing agent and turns into the alkoxy sulfate radical. The alkoxy sulfate radical can undergo electron transfer from a suitable donor and turn into the sulfate butanediol product or can undergo hydrogen abstraction by an oxygen molecule and turn into the hydroxybutanone product. Besides, the alkoxy sulfate radical can participate in several fragmentation processes to give smaller molecules like glycolic sulfates, acetic acid, formic acid, and methanol.
Hansel et al. [21,271] proposed the chemical mechanisms of aqueous-phase transformation of MeJa and MeSa. For MeJa, the mechanism started with the addition of OH radical to a double C=C bond producing four isomeric hydroxyalkyl radicals (Scheme 47). Then, an O2 molecule is added to each isomer to give four isomers of a hydroperoxide radical. Each hydroperoxide isomer could either react with other peroxy radicals present to give alcohols, carbonyls, and organic peroxides, or react with HO2 radicals and lose O2 to produce hydroyperoxides. MeSa could react with OH radicals by two distinct paths. They start with the OH addition to the aromatic ring and the phenolic hydrogen abstraction by OH (Scheme 48). In the first path, 1,2- and 1,4-dihydroxycyclohexadienyl radicals formed, which reacted with molecular oxygen to give the peroxyl radicals. The peroxyl radical decomposed by elimination of an HO2 radical to give dihydroxybenzene. The latter could similarly react with OH and O2 to afford trihydroxybenzene products. Dihydroxycyclohexadienyl radicals could also lose an H2O molecule and form an H-adduct radical, converting to phenoxyl radical. The two radicals can recombine to form dimers. In the hydrogen abstraction path, the phenoxyl radical forms directly from MeSa. Trihydroxybenzene products can from peroxy radicals via an alternative pathway (Scheme 49).
An interesting concept was the pinacol rearrangement of protonated MBO, which led to carbonyl derivatives (Scheme 50) [132].
The ozonolysis of GLV has been studied in multiphase chamber experiments but not in the aqueous phase alone [48,92,198,279]. Hydroperoxides formed in the gas phase possibly partitioned to the particle aqueous-phase, and participated in acid-catalyzed sulfation (Scheme 45) [279]. 3-Hydroxypropanal, produced by the decomposition of primary (Z)-3-hexen-1-ol ozonide, reactively partitioned to the particle phase. Therein, it plausibly entered several oligomerization reactions to give higher molecular weight products (Scheme 51) [65,198]. A similar mechanism is valid for the particle-phase oligomerization of 3-oxopropyl acetate and propanal (Scheme 52) [92]. Both compounds originated from the gas-phase ozonolysis of (Z)-3-hexenyl acetate. The formation of several particle-phase products from 1-octen-3-ol through Criegee intermediates was proposed based on AMS analyses of SOA (Scheme 53) [48].

7. Atmospheric Impact of GLV

Holopainen et al. [72] provided a very recent concept review on the role of plant volatiles in the atmosphere-biosphere relations Like F.W. Went in his visionary approach from the 1960s [304,305], they elaborated on a framework linking plant emission with atmospheric processes and formation of soil components. However, that paper presented a general idea and did not go into a detailed treatment.

7.1. Atmospheric Lifetimes of GLV

An atmospheric lifetime of a given compound C, which decays in a gas-phase bimolecular reaction with an atmospheric oxidant X, is estimated based on the assumption that the rate of C decay follows Equation (22):
d   C d t = k X C
We assume that the concentration of X is constant, so Equation (22) becomes the explicitly integrable pseudo-first-order equation:
l n C C 0 = k t
For the simplicity of comparison, we assume that the atmospheric lifetime is the time ratio [C]/[C]0 that will be equal to e. After rearrangement, we obtain:
t a t m o s = 1 k X
By analogy, the atmospheric lifetime of a compound due to the gas-phase photolysis is:
t a t m o s = 1 j
where j is the rate constant of the photolytic decay of this compound.
The atmospheric lifetimes of GLV reported in the literature or calculated from reported rate constants are collected in Table 41.
Sarang et al. [163] analyzed the atmospheric significance of the aqueous-phase reactions of 1-penten-3-ol, (Z)-2-hexen-1-ol, and (E)-2-hexen-1-al with OH, NO3, and SO4 radicals. The aqueous-phase reactions reduced the atmospheric lifetimes of GLV at sufficiently high liquid water contents (LWC). The radical concentrations in the aqueous phase predicted by CAPRAM modeling [306] were: [OH] = 3.5 × 10−15–3 × 10−12 M, [NO3] = 1.9 × 10−15–8.6 × 10−14 M, [SO4] = 2.3 × 10−15–3.6 × 10−13 M. LWC ranged from 10−12 to 10−4 m3 m−3 (aerosol to storm). Henry’s equilibria governed the partitioning of reactants between the phases. Fluxes of 1-penten-3-ol and (Z)-2-hexen-1-ol removed by the gas-phase and aqueous-phase reactions were comparable only in urban and remote clouds with high liquid water content. (E)-2-hexenal was removed faster by gas-phase processing in all clouds. Generally, the aqueous-phase processes did not influence the GLV lifetimes in systems with LWC lower than 10−6 m3 m−3 that is most clouds and all aerosol systems. In systems with high LWC, like storms, the lifetime of 1-penten-3-ol decreased markedly due to the aqueous-phase reactions with OH radicals. Similarly, the lifetimes of all three GLV studied decreased due to the reactions with NO3. Aqueous-phase reactions of GLV with SO4 radicals could decrease the GLV lifetimes from years to hours at high LWC and radical concentrations. The authors estimated that the removal of GLV from the atmosphere by the aqueous-phase reaction with SO4 was faster than the removal by combined gas and aqueous phase reactions with OH or NO3 in clouds and rains of high LWC, provided those radicals were in sufficient proportion to OH and NO3 radicals. For 1-penten-3-ol the proportions were: [OH]/[SO4] ≤ 0.16, and [NO3]/[SO4] < 6.5; for (Z)-2-hexen-1-ol–[OH]/[SO4] ≤ 0.40, and [NO3]/[SO4] < 3; and for (E)-2-hexen-1-al–[OH]/[SO4] ≤ 0.11, and [NO3]/[SO4] < 1.

7.2. SOA Potential of GLV

There is strong evidence that GLV contribute to the formation of ambient SOA (Section 6). Several smog-chamber experiments (Section 6.1) demonstrated the SOA formation from individual GLV, GLV mixtures, and plant emissions. To date, several compounds identified in ambient aerosol samples are known to form by reactions of GLV with atmospheric oxidants (Section 6.2).
Hamilton [65] estimated that the global SOA production from GLV (namely (Z)-3-hexen-1-ol and (Z)-3-hexenyl acetate) was 1–5 TgC yr−1. The calculation based on SOA mass yields from the OH photooxidation of isoprene and each GLV in the smog chamber experiments (1.2%, 3.1%, and 0.93%, resp., Section 6.1) scaled up to 3%, 7.5%, and 2.25%. The scaled GLV yields were totaled and applied to the yearly emission of a group of bVOC, including hexenal, hexenol, hexenyl acetate, and six non-GLV compounds estimated as 10–50 TgC yr−1 [45]. However, it is not clear whether that range regarded the whole group or each of its members. Thus, the estimated production of 1–5 TgC yr−1 of SOA from GLV is somewhat imprecise and uncertain. Furthermore, the chamber yields reflected the masses of SOA carbon obtained from the parent compounds. They cannot convert any amount of a parent compound to the amount of derived SOA if the molecular composition of that compound is unknown.
Better estimates of global annual emissions are available for (Z)-3-hexenal (4.9 Tg yr−1), (Z)-3-hexenol (2.9 Tg yr−1), and MBO (2.2 Tg yr−1)—claimed to be the most abundant GLV (Section 1, [50]). Yields of SOA from OH-initiated processing of (Z)-3-hexen-1-ol in smog chambers ranged from 2% [279] to the scaled-up value of 7.5% [65]. The corresponding global annual formation of SOA from (Z)-3-hexen-1-ol ranges from 0.058 to 0.218 Tg yr−1. The maximum SOA yield from the smog-chamber ozonolysis of (Z)-3-hexenol was 5.2% [279], which gives 0.151 Tg yr−1 of SOA globally. Smog-chamber yields of SOA from 2-methyl-3-buten-2-ol were 0.7% for OH reactions and 0.3–1.8% for ozonolysis. Respective global annual production of SOA is 0.015 Tg yr−1 and 0.007–0.040 Tg yr−1, respectively. We can assume that the chamber SOA yields from the oxidation of (Z)-3-hexanal are the same as for (Z)-3-hexen-1-ol. Then, the atmospheric SOA yields from(Z)-3-hexanal are 0.098–0.368 Tg yr−1 from the OH reactions and 0.255 Tg yr−1 from the ozonolysis. Thus, the estimated overall SOA yields from (Z)-3-hexen-1-ol, MBO, and (Z)-3-hexenal would be 0.58–1.05 Tg yr−1.
Besides overall global estimates, one can calculate SOA emission from individual sources. The annual SOA production from C6 GLV alcohols and aldehydes emitted by cut grass in Sydney and Melbourne was 0.1–0.2 Gg yr−1, based on the emission estimates [19] (Section 2.3) and assumed 5% SOA yield. The annual global production of 940 μg m−2 yr−1 SOA from grass mowing [92] multiplied by 55.5 × 106 km2 of the total area of world grassland [307] translates into 52 Gg yr−1 of SOA.
Switchgrass, grown to produce biofuels, produces 0.7 kg ha−1 yr−1 of 1-penten-3-ol, hexenols, and hexanals during growth and harvest (based on [88], Table 2). If the SOA yield from GLV is assumed at 5% and the total cultivation area is 9.5 × 106 km2 (data for 16 major countries in 2013, [308]), the total SOA input is 33 Gg yr−1. That number can increase because the available marginal land suitable for switchgrass cultivation amounts to 22.3 × 106 km2 [309].
Estimating SOA production from GLV based on SOA yields from smog-chamber experiments and GLV emission is somewhat approximate. The formation and growth of aerosol in the atmosphere depends on many factors that differ between the chamber and atmospheric conditions like the concentration of reactants, time of reactions, competition by other reactants, and influence of meteorological phenomena.
The global yield of SOA should increase due to the aqueous-phase processing of GLV. Based on the calculated values of the Henry’s constant, Hansel et al. [21] concluded that the intermediate and final products of MeJa and MeSa reactions with OH radicals were better soluble in water than the parent compounds and were plausible candidates for components of aqueous SOA.
Richards-Henderson et al. [22] estimated that for typical atmospheric conditions, the reactions of (Z)-3-hexen-1-ol, (Z)-3-hexenyl acetate, MeSa, MeJa, and MBO with OH radicals in the aqueous phase produced SOA, albeit about 15 times less than the gas phase reactions. The primary reason was low solubility of GLV in water. One can expect this ratio to change if the GLV solubility increases due to factors such as ionic strength or the presence of other organic compounds.
Richards-Henderson et al. [265] estimated the aqueous-phase reactions of singlet oxygen 1O2 * with (Z)-3-hexen-1-ol, cis-3-hexenyl acetate, MeSa, MeJa, and MBO were relatively insignificant atmospherically. In contrast, the reactions of 3DMB * with (Z)-3-hexenyl acetate, MeSa, and MeJa produced a more aqueous SOA. They showed that the aqueous-phase reactions of MeJa with OH and 3DMB * are the major conversion paths creating SOA, while other GLV convert to SOA mostly in the gas-phase reactions with OH and O3.
Hansel et al. [21] found that MeJa and MeSa oxidation products in the aqueous phase were much less volatile than the parent compounds, so they were likely to remain in the condensed phases as SOA components.

8. Conclusions

This review proves the substantial interest in Green Leaf Volatiles chemistry among atmospheric scientists. GLV should remain mainly in the atmospheric gas phase because of limited solubility in water. Their aqueous-phase reactions may be significant in atmospheric systems of high liquid water contents. GLV processing at the interfaces, both air-water, and air-particle, is practically a terra incognita. A few papers indicate it may play a role [156,157,269].
The SOA formation from GLV has been proved and warrants further research. Although the previous estimates of SOA production were overestimated (1–5 TgC yr−1 [65]), the values estimated in this work are still substantial and range from 0.58 to 1.05 Tg yr−1 from (Z)-hexen-1-ol, (Z)-3-hexenal and MBO jointly, 33 Gg yr−1 from switch grass cultivation for biofuels, or 52 Gg yr−1 from grass mowing. All calculations utilize SOA yields determined in smog-chamber experiments, which consequently deserve research attention.
The GLV transformation in the atmosphere may be minor globally, compared to major players like isoprene or monoterpenes (mostly α-pinene). The main reason is far smaller GLV emission fluxes. However, the GLV influence on the air quality may be very significant under local scenarios, including high emissions of GLV, like harvesting, switchgrass cultivation, or lawn mowing. It may increase if fumigation of plants with GLV is widely introduced to agriculture, horticulture, and forestry.
Further research should focus on better estimation of SOA yields in smog-chamber experiments, including the influence of relative humidity and acidity of particles, better introduction of GLV to air quality models, at least for local scenarios, and better estimation of GLV emissions for the latter purpose. Aqueous-phase studies should pay more attention to interfacial processes and product characterization. More experimental data on solubility and air-water partitioning of GLV would certainly be welcomed.
The health effects of GLV and GLV atmospheric transformation products have not been studied and warrant further research via abiotic and biotic experiments. The results may be significant to social groups exposed to high GLV emissions like farmers and residents of houses with grass yards.

Author Contributions

Conceptualization, K.J.R.; data mining, K.S. and K.J.R.; writing, K.S., K.J.R., and R.S. All authors have read and agreed to the published version of the manuscript.

Funding

Kumar Sarang received funding for the scientific work from the European Commission Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreements number 711859, and from the financial resources for science in the years 2017–2021 awarded by the Polish Ministry of Science and Higher Education for the implementation of an international cofinanced project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the invitation from Atmosphere Editors to contribute to the Special Issue “Plant-Derived Volatiles and Their Contribution to Secondary Organic Aerosol”.

Conflicts of Interest

The authors declar no conflict of interest.

Abbreviations

AATalcohol acyltransferase
ADHalcohol dehydrogenase
AMSaerosol mass spectrometry
AOCallene oxide cyclase
AOSallene oxide synthase
Aquaqueous
BVOCbiogenic volatile organic compounds
CEAScavity-enhanced absorption spectroscopy
CRACsmog chamber at the University of Cork
DFTdensity functional theory
DHIPdihydroxoisopropanol
DMAPPdimethylallyl pyrophosphate (dimethylallyl diphosphate)
DWdry weight (mass)
EIelectron ionization (in mass spectrometry)
EUPHOREenvironmental smog chamber in Valencia, Spain
EVKethyl vinyl ketone (1-penten-3-one)
FIDflame ionization detector (in gas chromatography)
FTIRFourier transform infrared spectroscopy
GCgas chromatography
GLVgreen leaf volatiles
HMWhigher molecular weight
JAjasmonic acid
LCliquid chromatography
LDIlaser desorption ionization (in mass spectrometry)
LFERlinear free energy relation
LISAsmog chamber at the Laboratoire Interuniversitaire des Systèmes Atmosphériques in Paris
LOXlipoxygenase
LWCliquid water contents, m3 (water) m−3 (air)
MBO2-methyl-3-buten-2-ol
MEGANModel of Emission of Gases and Aerosols from Nature
MeJamethyl jasmonate
MeSamethyl salicylate
MSmass spectrometry
MWmolecular weight
NIRnear infra-red (spectroscopy)
PARphotosynthetically active radiation
PIphotoionization detector
PLP-LIFpulsed laser photolysis—laser-induced fluorescence
PLP-RFpulse laser photolysis—resonance fluorescence
PMparticulate matter, ambient aerosol
RHrelative humidity (%)
RRKMRice–Ramsperger–Kassel–Marcus theory
SARstructure-activity relation
SMPSscanning mobility particle sizer
SOAsecondary organic aerosol
SPMEsolid-phase microextraction
SRRstructure-reactivity relation
TOFtime of flight analyzer (in mass spectrometry)
UVultraviolet
VOCvolatile organic compounds
yr, yrsyear, years

References

  1. Matsui, K.; Koeduka, T. Green leaf volatiles in plant signaling and response. In Lipids in Plant and Algae Development; Nakamura, Y., LiBeisson, Y., Eds.; Springer: New York, NY, USA, 2016; Volume 86, pp. 427–443. [Google Scholar]
  2. Dong, F.; Fu, X.M.; Watanabe, N.; Su, X.G.; Yang, Z.Y. Recent Advances in the Emission and Functions of Plant Vegetative Volatiles. Molecules 2016, 21, 124. [Google Scholar] [CrossRef] [PubMed]
  3. Holopainen, J.K. Can forest trees compensate for stress-generated growth losses by induced production of volatile compounds? Tree Physiology 2011, 31, 1356–1377. [Google Scholar] [CrossRef]
  4. Baldwin, I.T.; Halitschke, R.; Paschold, A.; von Dahl, C.C.; Preston, C.A. Volatile Signaling in Plant-Plant Interactions: “Talking Trees” in the Genomics Era. Science 2006, 311, 812–815. [Google Scholar] [CrossRef] [Green Version]
  5. Arimura, G.; Pearse, I.S. Chapter One—From the Lab Bench to the Forest: Ecology and Defence Mechanisms of Volatile-Mediated ‘Talking Trees’. In Advances in Botanical Research; Becard, G., Ed.; Academic Press: Cambridge, MA, USA, 2017; Volume 82, pp. 3–17. [Google Scholar]
  6. Farmer, E.E. Surface-to-air signals. Nature 2001, 411, 854–856. [Google Scholar] [CrossRef]
  7. Shulaev, V.; Silverman, P.; Raskin, I. Airborne signalling by methyl salicylate in plant pathogen resistance. Nature 1997, 385, 718–721. [Google Scholar] [CrossRef]
  8. Loreto, F.; Schnitzler, J.-P. Abiotic stresses and induced BVOCs. Trends Plant Sci. 2010, 15, 154–166. [Google Scholar] [CrossRef] [PubMed]
  9. Shen, J.Y.; Tieman, D.; Jones, J.B.; Taylor, M.G.; Schmelz, E.; Huffaker, A.; Bies, D.; Chen, K.S.; Klee, H.J. A 13-lipoxygenase, TomloxC, is essential for synthesis of C5 flavour volatiles in tomato. J. Exp. Bot. 2014, 65, 419–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Hatanaka, A. The biogeneration of green odor by green leaves. Phytochemistry 1993, 34, 1201–1218. [Google Scholar] [CrossRef]
  11. Matsui, K.; Sugimoto, K.; Mano, J.; Ozawa, R.; Takabayashi, J. Differential Metabolisms of Green Leaf Volatiles in Injured and Intact Parts of a Wounded Leaf Meet Distinct Ecophysiological Requirements. PLoS ONE 2012, 7, e10036433. [Google Scholar] [CrossRef]
  12. de Gouw, J.A.; Howard, C.J.; Custer, T.G.; Fall, R. Emissions of volatile organic compounds from cut grass and clover are enhanced during the drying process. Geophys. Res. Lett. 1999, 26, 811–814. [Google Scholar] [CrossRef]
  13. Jardine, K.; Barron-Gafford, G.A.; Norman, J.P.; Abrell, L.; Monson, R.K.; Meyers, K.T.; Pavao-Zuckerman, M.; Dontsova, K.; Kleist, E.; Werner, C.; et al. Green leaf volatiles and oxygenated metabolite emission bursts from mesquite branches following light-dark transitions. Photosynth. Res. 2012, 113, 321–333. [Google Scholar] [CrossRef]
  14. Loreto, F.; Barta, C.; Brilli, F.; Nogues, I. On the induction of volatile organic compound emissions by plants as consequence of wounding or fluctuations of light and temperature. Plant Cell Environ. 2006, 29, 1820–1828. [Google Scholar] [CrossRef]
  15. Brilli, F.; Ruuskanen, T.M.; Schnitzhofer, R.; Muller, M.; Breitenlechner, M.; Bittner, V.; Wohlfahrt, G.; Loreto, F.; Hansel, A. Detection of Plant Volatiles after Leaf Wounding and Darkening by Proton Transfer Reaction "Time-of-Flight" Mass Spectrometry (PTR-TOF). PLoS ONE 2011, 6, e0020419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Jardine, K.J.; Chambers, J.Q.; Holm, J.; Jardine, A.B.; Fontes, C.G.; Zorzanelli, R.F.; Meyers, K.T.; de Souza, V.F.; Garcia, S.; Gimenez, B.O.; et al. Green leaf volatile emissions during high temperature and drought stress in a Central Amazon rainforest. Plants 2015, 4, 678. [Google Scholar] [CrossRef] [Green Version]
  17. Karl, T.; Harren, F.; Warneke, C.; de Gouw, J.; Grayless, C.; Fall, R. Senescing grass crops as regional sources of reactive volatile organic compounds. J. Geophys. Res. Atmos. 2005, 110, 12. [Google Scholar] [CrossRef] [Green Version]
  18. Kirstine, W.; Galbally, I.; Ye, Y.R.; Hooper, M. Emissions of volatile organic compounds (primarily oxygenated species) from pasture. J. Geophys. Res. Atmos. 1998, 103, 10605–10619. [Google Scholar] [CrossRef]
  19. Kirstine, W.V.; Galbally, I.E. A simple model for estimating emissions of volatile organic compounds from grass and cut grass in urban airsheds and its application to two Australian cities. J. Air Waste Manag. Assoc. 2004, 54, 1299–1311. [Google Scholar] [CrossRef]
  20. Scala, A.; Allmann, S.; Mirabella, R.; Haring, M.; Schuurink, R. Green Leaf Volatiles: A Plant’s Multifunctional Weapon against Herbivores and Pathogens. Int. J. Mol. Sci. 2013, 14, 17781. [Google Scholar] [CrossRef] [Green Version]
  21. Hansel, A.K.; Ehrenhauser, F.S.; Richards-Henderson, N.K.; Anastasio, C.; Valsaraj, K.T. Aqueous-phase oxidation of green leaf volatiles by hydroxyl radical as a source of SOA: Product identification from methyl jasmonate and methyl salicylate oxidation. Atmos. Environ. 2015, 102, 43–51. [Google Scholar] [CrossRef]
  22. Richards-Henderson, N.K.; Hansel, A.K.; Valsaraj, K.T.; Anastasio, C. Aqueous oxidation of green leaf volatiles by hydroxyl radical as a source of SOA: Kinetics and SOA yields. Atmos. Environ. 2014, 95, 105–112. [Google Scholar] [CrossRef]
  23. Maffei, M.E. Sites of synthesis, biochemistry and functional role of plant volatiles. S. Afr. J. Bot. 2010, 76, 612–631. [Google Scholar] [CrossRef] [Green Version]
  24. Fisher, A.J.; Baker, B.M.; Greenberg, J.P.; Fall, R. Enzymatic Synthesis of Methylbutenol from Dimethylallyl Diphosphate in Needles of Pinus sabiniana. Arch. Biochem. Biophys. 2000, 383, 128–134. [Google Scholar] [CrossRef]
  25. Guenther, A.; Geron, C.; Pierce, T.; Lamb, B.; Harley, P.; Fall, R. Natural emissions of non-methane volatile organic compounds; carbon monoxide, and oxides of nitrogen from North America. Atmos. Environ. 2000, 34, 2205–2230. [Google Scholar] [CrossRef] [Green Version]
  26. Mochizuki, S.; Sugimoto, K.; Koeduka, T.; Matsui, K. Arabidopsis lipoxygenase 2 is essential for formation of green leaf volatiles and five-carbon volatiles. FEBS Lett. 2016, 590, 1017–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Fisher, A.J.; Grimes, H.D.; Fall, R. The biochemical origin of pentenol emissions from wounded leaves. Phytochemistry 2003, 62, 159–163. [Google Scholar] [CrossRef]
  28. Mayland, H.F.; Flath, R.A.; Shewmaker, G.E. Volatiles from Fresh and Air-Dried Vegetative Tissues of Tall Fescue (Festuca arundinacea Schreb.): Relationship to Cattle Preference. J. Agric. Food Chem. 1997, 45, 2204–2210. [Google Scholar] [CrossRef]
  29. Fall, R.; Karl, T.; Jordon, A.; Lindinger, W. Biogenic C5VOCs: Release from leaves after freeze-thaw wounding and occurrence in air at a high mountain observatory. Atmos. Environ. 2001, 35, 3905–3916. [Google Scholar] [CrossRef]
  30. López-Gresa, M.P.; Lisón, P.; Campos, L.; Rodrigo, I.; Rambla, J.L.; Granell, A.; Conejero, V.; Bellés, J.M. A Non-targeted Metabolomics Approach Unravels the VOCs Associated with the Tomato Immune Response against Pseudomonas syringae. Front. Plant Sci. 2017, 8, 1188. [Google Scholar] [CrossRef] [Green Version]
  31. Connor, E.C.; Rott, A.S.; Zeder, M.; Juttner, F.; Dorn, S. (13)C-labelling patterns of green leaf volatiles indicating different dynamics of precursors in Brassica leaves. Phytochemistry 2008, 69, 1304–1312. [Google Scholar] [CrossRef] [PubMed]
  32. Matsui, K. Green leaf volatiles: Hydroperoxide lyase pathway of oxylipin metabolism. Curr. Opin. Plant Biol. 2006, 9, 274–280. [Google Scholar] [CrossRef]
  33. Fall, R.; Karl, T.; Hansel, A.; Jordan, A.; Lindinger, W. Volatile organic compounds emitted after leaf wounding: On-line analysis by proton-transfer-reaction mass spectrometry. J. Geophys. Res. Atmos. 1999, 104, 15963–15974. [Google Scholar] [CrossRef]
  34. Engelberth, J.; Alborn, H.T.; Schmelz, E.A.; Tumlinson, J.H. Airborne signals prime plants against insect herbivore attack. Proc. Natl. Acad. Sci. USA 2004, 101, 1781–1785. [Google Scholar] [CrossRef] [Green Version]
  35. Croft, K.P.C.; Juttner, F.; Slusarenko, A.J. Volatile products of the lipoxygenase pathway evolved from Phaseolus-vulgaris (L.) leaves inoculated with Pseudomonas syringae pv phaseolicola. Plant Physiol. 1993, 101, 13–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Fukui, Y.; Doskey, P.V. Identification of nonmethane organic compound emissions from grassland vegetation. Atmos. Environ. 2000, 34, 2947–2956. [Google Scholar] [CrossRef]
  37. Rose, U.; Manukian, A.; Heath, R.R.; Tumlinson, J.H. Volatile Semiochemicals Released from Undamaged Cotton Leaves (A Systemic Response of Living Plants to Caterpillar Damage). Plant Physiol. 1996, 111, 487–495. [Google Scholar] [CrossRef] [Green Version]
  38. Matsui, K.; Minami, A.; Hornung, E.; Shibata, H.; Kishimoto, K.; Ahnert, V.; Kindl, H.; Kajiwara, T.; Feussner, I. Biosynthesis of fatty acid derived aldehydes is induced upon mechanical wounding and its products show fungicidal activities in cucumber. Phytochemistry 2006, 67, 649–657. [Google Scholar] [CrossRef] [PubMed]
  39. Matsui, K.; Kurishita, S.; Hisamitsu, A.; Kajiwara, T. A lipid-hydrolysing activity involved in hexenal formation. Biochem. Soc. Trans. 2000, 28, 857–860. [Google Scholar] [CrossRef]
  40. Peterson, D.L.; Böröczky, K.; Tumlinson, J.; Cipollini, D. Ecological fitting: Chemical profiles of plant hosts provide insights on selection cues and preferences for a major buprestid pest. Phytochemistry 2020, 176, 112397. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, Q.-H.; Birgersson, G.; Zhu, J.; Löfstedt, C.; Löfqvist, J.; Schlyter, F. Leaf Volatiles from Nonhost Deciduous Trees: Variation by Tree Species, Season and Temperature, and Electrophysiological Activity in Ips typographus. J. Chem. Ecol. 1999, 25, 1923–1943. [Google Scholar] [CrossRef]
  42. Gaquerel, E.; Weinhold, A.; Baldwin, I.T. Molecular Interactions between the Specialist Herbivore Manduca sexta (Lepidoptera, Sphigidae) and Its Natural Host Nicotiana attenuata. VIII. An Unbiased GCxGC-ToFMS Analysis of the Plant’s Elicited Volatile Emissions. Plant Physiol. 2009, 149, 1408–1423. [Google Scholar] [CrossRef] [Green Version]
  43. Kessler, A.; Baldwin, I.T. Defensive function of herbivore-induced plant volatile emissions in nature. Science 2001, 291, 2141–2144. [Google Scholar] [CrossRef] [PubMed]
  44. Blande, J.D.; Korjus, M.; Holopainen, J.K. Foliar methyl salicylate emissions indicate prolonged aphid infestation on silver birch and black alder. Tree Physiol. 2010, 30, 404–416. [Google Scholar] [CrossRef] [PubMed]
  45. Wiedinmyer, C.; Guenther, A.; Harley, P.; Hewitt, N.; Geron, C.; Artaxo, P.; Steinbrecher, R.; Rasmussen, R. Global organic emissions from vegetation. In Emissions of Atmospheric Trace Compound; Granier, C., Artaxo, P., Reeves, C.E., Eds.; Springer: Dordrecht, The Netherlands, 2004; pp. 115–170. [Google Scholar] [CrossRef]
  46. Karl, T.G.; Spirig, C.; Rinne, J.; Stroud, C.; Prevost, P.; Greenberg, J.; Fall, R.; Guenther, A. Virtual disjunct eddy covariance measurements of organic compound fluxes from a subalpine forest using proton transfer reaction mass spectrometry. Atmos. Chem. Phys. 2002, 2, 279–291. [Google Scholar] [CrossRef] [Green Version]
  47. Geron, C.D.; Daly, R.W.; Arnts, R.R.; Guenther, A.B.; Mowry, F.L. Canopy level emissions of 2-methyl-3-buten-2-ol, monoterpenes, and sesquiterpenes from an experimental Pinus taeda plantation. Sci. Total Environ. 2016, 565, 730–741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Fischer, K.B.; Gold, C.S.; Harvey, R.M.; Petrucci, A.N.; Petrucci, G.A. Ozonolysis Chemistry and Phase Behavior of 1-Octen-3-ol-Derived Secondary Organic Aerosol. ACS Earth Space Chem. 2020, 4, 1298–1308. [Google Scholar] [CrossRef]
  49. Creelman, R.A.; Mullet, J.E. Biosynthesis and action of jasmonates in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1997, 48, 355–381. [Google Scholar] [CrossRef] [Green Version]
  50. Guenther, A.B.; Jiang, X.; Heald, C.L.; Sakulyanontvittaya, T.; Duhl, T.; Emmons, L.K.; Wang, X. The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1): An extended and updated framework for modeling biogenic emissions. Geosci. Model Dev. 2012, 5, 1471–1492. [Google Scholar] [CrossRef] [Green Version]
  51. Sindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Müller, J.F.; Kuhn, U.; Stefani, P.; Knorr, W. Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 years. Atmos. Chem. Phys. 2014, 14, 9317–9341. [Google Scholar] [CrossRef] [Green Version]
  52. Su, Q.; Yang, F.; Zhang, Q.; Tong, H.; Hu, Y.; Zhang, X.; Xie, W.; Wang, S.; Wu, Q.; Zhang, Y. Defence priming in tomato by the green leaf volatile (Z)-3-hexenol reduces whitefly transmission of a plant virus. Plant Cell Environ. 2020, 43, 2797–2811. [Google Scholar] [CrossRef] [PubMed]
  53. Guo, R.-Z.; Yan, H.-Y.; Li, X.-X.; Zou, X.-X.; Zhang, X.-J.; Yu, X.-N.; Ci, D.-W.; Wang, Y.-F.; Si, T. Green leaf volatile (Z)-3-hexeny-1-yl acetate reduces salt stress in peanut by affecting photosynthesis and cellular redox homeostasis. Physiol. Plant. 2020, 170, 75–92. [Google Scholar] [CrossRef]
  54. Wakai, J.; Kusama, S.; Nakajima, K.; Kawai, S.; Okumura, Y.; Shiojiri, K. Effects of trans-2-hexenal and cis-3-hexenal on post-harvest strawberry. Sci. Rep. 2019, 9, 10112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Cofer, T.M.; Engelberth, M.; Engelberth, J. Green leaf volatiles protect maize (Zea mays) seedlings against damage from cold stress. Plant Cell Environ. 2018, 41, 1673–1682. [Google Scholar] [CrossRef] [PubMed]
  56. Payá, C.; López-Gresa, M.P.; Intrigliolo, D.S.; Rodrigo, I.; Bellés, J.M.; Lisón, P. (Z)-3-Hexenyl Butyrate Induces Stomata Closure and Ripening in Vitis vinifera. Agronomy 2020, 10, 1122. [Google Scholar] [CrossRef]
  57. Jiménez, E.; Lanza, B.; Antiñolo, M.; Albaladejo, J. Photooxidation of leaf-wound oxygenated compounds, 1-penten-3-ol, (Z)-3-hexen-1-ol, and 1-penten-3-one, initiated by OH radicals and sunlight. Environ. Sci. Technol. 2009, 43, 1831–1837. [Google Scholar] [CrossRef]
  58. Yli-Pirilä, P.; Copolovici, L.; Kännaste, A.; Noe, S.; Blande, J.D.; Mikkonen, S.; Klemola, T.; Pulkkinen, J.; Virtanen, A.; Laaksonen, A.; et al. Herbivory by an Outbreaking Moth Increases Emissions of Biogenic Volatiles and Leads to Enhanced Secondary Organic Aerosol Formation Capacity. Environ. Sci. Technol. 2016, 50, 11501–11510. [Google Scholar] [CrossRef] [Green Version]
  59. Tsigaridis, K.; Daskalakis, N.; Kanakidou, M.; Adams, P.J.; Artaxo, P.; Bahadur, R.; Balkanski, Y.; Bauer, S.E.; Bellouin, N.; Benedetti, A.; et al. The AeroCom evaluation and intercomparison of organic aerosol in global models. Atmos. Chem. Phys. 2014, 14, 10845–10895. [Google Scholar] [CrossRef] [Green Version]
  60. Khan, M.A.H.; Jenkin, M.E.; Foulds, A.; Derwent, R.G.; Percival, C.J.; Shallcross, D.E. A modeling study of secondary organic aerosol formation from sesquiterpenes using the STOCHEM global chemistry and transport model. J. Geophys. Res. Atmos. 2017, 122, 4426–4439. [Google Scholar] [CrossRef]
  61. Hodzic, A.; Kasibhatla, P.S.; Jo, D.S.; Cappa, C.D.; Jimenez, J.L.; Madronich, S.; Park, R.J. Rethinking the global secondary organic aerosol (SOA) budget: Stronger production, faster removal, shorter lifetime. Atmos. Chem. Phys. 2016, 16, 7917–7941. [Google Scholar] [CrossRef] [Green Version]
  62. Carlton, A.G.; Wiedinmyer, C.; Kroll, J.H. A review of secondary organic aerosol (SOA) formation from isoprene. Atmos. Chem. Phys. 2009, 9, 4987–5005. [Google Scholar] [CrossRef] [Green Version]
  63. Lim, H.-J.; Carlton, A.G.; Turpin, B.J. Isoprene Forms Secondary Organic Aerosol through Cloud Processing: Model Simulations. Environ. Sci. Technol. 2005, 39, 4441–4446. [Google Scholar] [CrossRef]
  64. Lin, G.; Penner, J.E.; Zhou, C. How will SOA change in the future? Geophys. Res. Lett. 2016, 43, 1718–1726. [Google Scholar] [CrossRef] [Green Version]
  65. Hamilton, J.F.; Lewis, A.C.; Carey, T.J.; Wenger, J.C.; Garcia, E.B.I.; Munoz, A. Reactive oxidation products promote secondary organic aerosol formation from green leaf volatiles. Atmos. Chem. Phys. 2009, 9, 3815–3823. [Google Scholar] [CrossRef] [Green Version]
  66. Dicke, M.; Loreto, F. Induced plant volatiles: From genes to climate change. Trends Plant Sci. 2010, 15, 115–117. [Google Scholar] [CrossRef] [PubMed]
  67. Peñuelas, J.; Staudt, M. BVOCs and global change. Trends Plant Sci. 2010, 15, 133–144. [Google Scholar] [CrossRef] [PubMed]
  68. Holopainen, J.K.; Gershenzon, J. Multiple stress factors and the emission of plant VOCs. Trends Plant Sci. 2010, 15, 176–184. [Google Scholar] [CrossRef]
  69. Noziere, 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, 3919–3983. [Google Scholar] [CrossRef]
  70. Herrmann, H.; Schaefer, T.; Tilgner, A.; Styler, S.A.; Weller, C.; Teich, M.; Otto, T. Tropospheric aqueous-phase chemistry: Kinetics, mechanisms, and its coupling to a changing gas phase. Chem. Rev. 2015, 115, 4259–4334. [Google Scholar] [CrossRef] [PubMed]
  71. Mellouki, A.; Wallington, T.J.; Chen, J. Atmospheric chemistry of oxygenated volatile organic compounds: Impacts on air quality and climate. Chem. Rev. 2015, 115, 3984–4014. [Google Scholar] [CrossRef] [PubMed]
  72. Holopainen, J.K.; Kivimäenpää, M.; Nizkorodov, S.A. Plant-derived Secondary Organic Material in the Air and Ecosystems. Trends Plant Sci. 2017, 22, 744–753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ameye, M.; Allmann, S.; Verwaeren, J.; Smagghe, G.; Haesaert, G.; Schuurink, R.C.; Audenaert, K. Green leaf volatile production by plants: A meta-analysis. New Phytol. 2018, 220, 666–683. [Google Scholar] [CrossRef]
  74. López-Gresa, M.P.; Payá, C.; Ozáez, M.; Rodrigo, I.; Conejero, V.; Klee, H.; Bellés, J.M.; Lisón, P. A New Role for Green Leaf Volatile Esters in Tomato Stomatal Defense Against Pseudomonas syringe pv. tomato. Front. Plant Sci. 2018, 9, 1855. [Google Scholar] [CrossRef] [PubMed]
  75. Kivimäenpää, M.; Babalola, A.B.; Joutsensaari, J.; Holopainen, J.K. Methyl Salicylate and Sesquiterpene Emissions Are Indicative for Aphid Infestation on Scots Pine. Forests 2020, 11, 573. [Google Scholar] [CrossRef]
  76. Niinemets, Ü.; Kännaste, A.; Copolovici, L. Quantitative patterns between plant volatile emissions induced by biotic stresses and the degree of damage. Front. Plant Sci. 2013, 4, 262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Mattiacci, L.; Rocca, B.A.; Scascighini, N.; D’Alessandro, M.; Hern, A.; Dorn, S. Systemically induced plant volatiles emitted at the time of “danger”. J. Chem. Ecol. 2001, 27, 2233–2252. [Google Scholar] [CrossRef] [PubMed]
  78. Engelberth, J. Primed to grow: A new role for green leaf volatiles in plant stress responses. Plant Signal. Behav. 2020, 15, 1701240. [Google Scholar] [CrossRef]
  79. Arimura, G.-I.; Kost, C.; Boland, W. Herbivore-induced, indirect plant defences. Biochim. Biophys. Acta (BBA) Mol. Cell Biol. Lipids 2005, 1734, 91–111. [Google Scholar] [CrossRef] [PubMed]
  80. Cui, H.; Wei, J.; Su, J.; Li, C.; Ge, F. Elevated O3 increases volatile organic compounds via jasmonic acid pathway that promote the preference of parasitoid Encarsia formosa for tomato plants. Plant Sci. 2016, 253, 243–250. [Google Scholar] [CrossRef] [PubMed]
  81. Pinto, D.M.; Nerg, A.M.; Holopainen, J.K. The role of ozone-reactive compounds, terpenes, and green leaf volatiles (GLVs), in the orientation of Cotesia plutellae. J. Chem. Ecol. 2007, 33, 2218–2228. [Google Scholar] [CrossRef] [PubMed]
  82. Hanley, M.E.; Shannon, R.W.R.; Lemoine, D.G.; Sandey, B.; Newland, P.L.; Poppy, G.M. Riding on the wind: Volatile compounds dictate selection of grassland seedlings by snails. Ann. Bot. 2018, 122, 1075–1083. [Google Scholar] [CrossRef] [PubMed]
  83. Dudareva, N.; Negre, F.; Nagegowda, D.A.; Orlova, I. Plant volatiles: Recent advances and future perspectives. Crit. Rev. Plant Sci. 2006, 25, 417–440. [Google Scholar] [CrossRef]
  84. Engelberth, J.; Engelberth, M. Variability in the Capacity to Produce Damage-Induced Aldehyde Green Leaf Volatiles among Different Plant Species Provides Novel Insights into Biosynthetic Diversity. Plants 2020, 9, 213. [Google Scholar] [CrossRef] [Green Version]
  85. Albaladejo, J.; Ballesteros, B.; Jiménez, E.; Martén, P.; Marténez, E. A PLP–LIF kinetic study of the atmospheric reactivity of a series of C4–C7 saturated and unsaturated aliphatic aldehydes with OH. Atmos. Environ. 2002, 36, 3231–3239. [Google Scholar] [CrossRef]
  86. König, G.; Brunda, M.; Puxbaum, H.; Hewitt, C.N.; Duckham, S.C.; Rudolph, J. Relative contribution of oxygenated hydrocarbons to the total biogenic VOC emissions of selected mid-European agricultural and natural plant species. Atmos. Environ. 1995, 29, 861–874. [Google Scholar] [CrossRef]
  87. García-Plazaola, J.I.; Portillo-Estrada, M.; Fernández-Marín, B.; Kännaste, A.; Niinemets, Ü. Emissions of carotenoid cleavage products upon heat shock and mechanical wounding from a foliose lichen. Environ. Exp. Bot. 2017, 133, 87–97. [Google Scholar] [CrossRef] [Green Version]
  88. Eller, A.S.D.; Sekimoto, K.; Gilman, J.B.; Kuster, W.C.; de Gouw, J.A.; Monson, R.K.; Graus, M.; Crespo, E.; Warneke, C.; Fall, R. Volatile organic compound emissions from switchgrass cultivars used as biofuel crops. Atmos. Environ. 2011, 45, 3333–3337. [Google Scholar] [CrossRef]
  89. Heiden, A.C.; Hoffmann, T.; Kahl, J.; Kley, D.; Klockow, D.; Langebartels, C.; Mehlhorn, H.; Sandermann, H., Jr.; Schraudner, M.; Schuh, G.; et al. Emission of volatile organic compounds from ozone-exposed plants. Ecol. Appl. 1999, 9, 1160–1167. [Google Scholar] [CrossRef]
  90. Helmig, D.; Klinger, L.F.; Guenther, A.; Vierling, L.; Geron, C.; Zimmerman, P. Biogenic volatile organic compound emissions (BVOCs) I. Identifications from three continental sites in the U.S. Chemosphere 1999, 38, 2163–2187. [Google Scholar] [CrossRef] [Green Version]
  91. Semiz, G.; Blande, J.D.; Heijari, J.; Işık, K.; Niinemets, Ü.; Holopainen, J.K. Manipulation of VOC emissions with methyl jasmonate and carrageenan in the evergreen conifer Pinus sylvestris and evergreen broadleaf Quercus ilex. Plant Biol. 2012, 14, 57–65. [Google Scholar] [CrossRef]
  92. Harvey, R.M.; Zahardis, J.; Petrucci, G.A. Establishing the contribution of lawn mowing to atmospheric aerosol levels in American suburbs. Atmos. Chem. Phys. 2014, 14, 797–812. [Google Scholar] [CrossRef] [Green Version]
  93. Giacomuzzi, V.; Cappellin, L.; Nones, S.; Khomenko, I.; Biasioli, F.; Knight, A.L.; Angeli, S. Diel rhythms in the volatile emission of apple and grape foliage. Phytochemistry 2017, 138, 104–115. [Google Scholar] [CrossRef]
  94. Fukui, Y.; Doskey, P.V. Air-surface exchange of nonmethane organic compounds at a grassland site: Seasonal variations and stressed emissions. J. Geophys. Res. Atmos. 1998, 103, 13153–13168. [Google Scholar] [CrossRef]
  95. Arey, J.; Winer, A.M.; Atkinson, R.; Aschmann, S.M.; Long, W.D.; Lynn Morrison, C. The emission of (Z)-3-hexen-1-ol, (Z)-3-hexenylacetate and other oxygenated hydrocarbons from agricultural plant species. Atmos. Environ. Part A Gen. Top. 1991, 25, 1063–1075. [Google Scholar] [CrossRef]
  96. Bamberger, I.; Hortnagl, L.; Schnitzhofer, R.; Graus, M.; Ruuskanen, T.M.; Muller, M.; Dunkl, J.; Wohlfahrt, G.; Hansel, A. BVOC fluxes above mountain grassland. Biogeosciences 2010, 7, 1413–1424. [Google Scholar] [CrossRef] [Green Version]
  97. Gray, D.W.; Lerdau, M.T.; Goldstein, A.H. Influences of temperature history, water stress, and needle age on methylbutenol emissions. Ecology 2003, 84, 765–776. [Google Scholar] [CrossRef]
  98. Harley, P.; Eller, A.; Guenther, A.; Monson, R.K. Observations and models of emissions of volatile terpenoid compounds from needles of ponderosa pine trees growing in situ: Control by light, temperature and stomatal conductance. Oecologia 2014, 176, 35–55. [Google Scholar] [CrossRef] [Green Version]
  99. Juráň, S.; Pallozzi, E.; Guidolotti, G.; Fares, S.; Šigut, L.; Calfapietra, C.; Alivernini, A.; Savi, F.; Večeřová, K.; Křůmal, K.; et al. Fluxes of biogenic volatile organic compounds above temperate Norway spruce forest of the Czech Republic. Agric. For. Meteorol. 2017, 232, 500–513. [Google Scholar] [CrossRef]
  100. Baker, B.; Guenther, A.; Greenberg, J.; Goldstein, A.; Fall, R. Canopy fluxes of 2-methyl-3-buten-2-ol over a ponderosa pine forest by relaxed eddy accumulation: Field data and model comparison. J. Geophys. Res. Atmos. 1999, 104, 26107–26114. [Google Scholar] [CrossRef] [Green Version]
  101. Schade, G.W.; Goldstein, A.H. Fluxes of oxygenated volatile organic compounds from a ponderosa pine plantation. J. Geophys. Res. Atmos. 2001, 106, 3111–3123. [Google Scholar] [CrossRef]
  102. Baker, B.; Guenther, A.; Greenberg, J.; Fall, R. Canopy Level Fluxes of 2-Methyl-3-buten-2-ol, Acetone, and Methanol by a Portable Relaxed Eddy Accumulation System. Environ. Sci. Technol. 2001, 35, 1701–1708. [Google Scholar] [CrossRef] [Green Version]
  103. Arnts, R.R.; Mowry, F.L.; Hampton, G.A. A high-frequency response relaxed eddy accumulation flux measurement system for sampling short-lived biogenic volatile organic compounds. J. Geophys. Res. Atmos. 2013, 118, 4860–4873. [Google Scholar] [CrossRef]
  104. Schade, G.W.; Goldstein, A.H.; Gray, D.W.; Lerdau, M.T. Canopy and leaf level 2-methyl-3-buten-2-ol fluxes from a ponderosa pine plantation. Atmos. Environ. 2000, 34, 3535–3544. [Google Scholar] [CrossRef]
  105. Harley, P.; Fridd-Stroud, V.; Greenberg, J.; Guenther, A.; Vasconcellos, P. Emission of 2-methyl-3-buten-2-ol by pines: A potentially large natural source of reactive carbon to the atmosphere. J. Geophys. Res. Atmos. 1998, 103, 25479–25486. [Google Scholar] [CrossRef] [Green Version]
  106. Karl, T.; Guenther, A.; Turnipseed, A.; Patton, E.G.; Jardine, K. Chemical sensing of plant stress at the ecosystem scale. Biogeosciences 2008, 5, 1287–1294. [Google Scholar] [CrossRef] [Green Version]
  107. Copolovici, L.; Kännaste, A.; Remmel, T.; Niinemets, Ü. Volatile organic compound emissions from Alnus glutinosa under interacting drought and herbivory stresses. Environ. Exp. Bot. 2014, 100, 55–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Brilli, F.; Hortnagl, L.; Bamberger, I.; Schnitzhofer, R.; Ruuskanen, T.M.; Hansel, A.; Loreto, F.; Wohlfahrt, G. Qualitative and quantitative characterization of volatile organic compound emissions from cut grass. Environ. Sci. Technol. 2012, 46, 3859–3865. [Google Scholar] [CrossRef]
  109. Karl, T.; Guenther, A.; Lindinger, C.; Jordan, A.; Fall, R.; Lindinger, W. Eddy covariance measurements of oxygenated volatile organic compound fluxes from crop harvesting using a redesigned proton-transfer-reaction mass spectrometer. J. Geophys. Res. Atmos. 2001, 106, 24157–24167. [Google Scholar] [CrossRef] [Green Version]
  110. Karl, T.; Guenther, A.; Jordan, A.; Fall, R.; Lindinger, W. Eddy covariance measurement of biogenic oxygenated VOC emissions from hay harvesting. Atmos. Environ. 2001, 35, 491–495. [Google Scholar] [CrossRef] [Green Version]
  111. de Carvalho, A.B.; Kato, M.; Rezende, M.M.; Pereira, P.; de Andrade, J.B. Determination of carbonyl compounds in the atmosphere of charcoal plants by HPLC and UV detection. J. Sep. Sci. 2008, 31, 1686–1693. [Google Scholar] [CrossRef] [PubMed]
  112. Karl, T.; Fall, R.; Jordan, A.; Lindinger, W. On-line analysis of reactive VOCs from urban lawn mowing. Environ. Sci. Technol. 2001, 35, 2926–2931. [Google Scholar] [CrossRef] [PubMed]
  113. Gentner, D.R.; Ormeno, E.; Fares, S.; Ford, T.B.; Weber, R.; Park, J.H.; Brioude, J.; Angevine, W.M.; Karlik, J.F.; Goldstein, A.H. Emissions of terpenoids, benzenoids, and other biogenic gas-phase organic compounds from agricultural crops and their potential implications for air quality. Atmos. Chem. Phys. 2014, 14, 5393–5413. [Google Scholar] [CrossRef] [Green Version]
  114. Goldan, P.D.; Kuster, W.C.; Fehsenfeld, F.C.; Montzka, S.A. The observation of a C5 alcohol emission in a North American pine forest. Geophys. Res. Lett. 1993, 20, 1039–1042. [Google Scholar] [CrossRef]
  115. 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, 3181–3210. [Google Scholar] [CrossRef] [Green Version]
  116. Heiden, A.C.; Kobel, K.; Langebartels, C.; Schuh-Thomas, G.; Wildt, J. Emissions of oxygenated volatile organic compounds from plants - part I: Emissions from lipoxygenase activity. J. Atmos. Chem. 2003, 45, 143–172. [Google Scholar] [CrossRef]
  117. US_EPA. Estimation Programs Interface Suite™ for Microsoft® Windows, v 4.11; United States Environmental Protection Agency: Washington, DC, USA, 2012. [Google Scholar]
  118. Hine, J.; Mookerjee, P.K. Structural effects on rates and equilibriums. XIX. Intrinsic hydrophilic character of organic compounds. Correlations in terms of structural contributions. J. Org. Chem. 1975, 40, 292–298. [Google Scholar] [CrossRef]
  119. Altschuh, J.; Bruggemann, R.; Santl, H.; Eichinger, G.; Piringer, O.G. Henry’s law constants for a diverse set of organic chemicals: Experimental determination and comparison of estimation methods. Chemosphere 1999, 39, 1871–1887. [Google Scholar] [CrossRef]
  120. Sander, R. Compilation of Henry’s law constants (version 4.0) for water as solvent. Atmos. Chem. Phys. 2015, 15, 4399–4981. [Google Scholar] [CrossRef] [Green Version]
  121. Staudinger, J.; Roberts, P.V. A critical review of Henry’s law constants for environmental applications. Crit. Rev. Environ. Sci. Technol. 1996, 26, 205–297. [Google Scholar] [CrossRef]
  122. Sangster, J. Octanol-water partition coefficients of simple organic compounds. J. Phys. Chem. Ref. Data 1989, 18, 1111–1229. [Google Scholar] [CrossRef]
  123. Abraham, M.H.; Le, J.; Acree, W.E.; Carr, P.W.; Dallas, A.J. The solubility of gases and vapours in dry octan-1-ol at 298 K. Chemosphere 2001, 44, 855–863. [Google Scholar] [CrossRef]
  124. Shunthirasingham, C.; Cao, X.S.; Lei, Y.D.; Wania, F. Large bubbles reduce the surface sorption artifact of the inert gas stripping method. J. Chem. Eng. Data 2013, 58, 792–797. [Google Scholar] [CrossRef]
  125. Hansch, C.; Quinlan, J.E.; Lawrence, G.L. Linear free-energy relationship between partition coefficients and aqueous solubility of organic liquids. J. Org. Chem. 1968, 33, 347. [Google Scholar] [CrossRef]
  126. Suzuki, T. Development of an automatic estimation system for both the partition-coefficient and aqueous solubility. J. Comput.-Aided Mol. Des. 1991, 5, 149–166. [Google Scholar] [CrossRef]
  127. Karl, T.; Yeretzian, C.; Jordan, A.; Lindinger, W. Dynamic measurements of partition coefficients using proton-transfer-reaction mass spectrometry (PTR–MS). Int. J. Mass Spectrom. 2003, 223–224, 383–395. [Google Scholar] [CrossRef]
  128. Sander, R. Compilation of Henry’s Law Constants for Inorganic and Organic Species of Potential Importance in Environmental Chemistry; Max-Planck Institute of Chemistry: Mainz, Germany, 1999; p. 107. [Google Scholar]
  129. Zhou, X.; Mopper, K. Apparent partition coefficients of 15 carbonyl compounds between air and seawater and between air and freshwater; implications for air-sea exchange. Environ. Sci. Technol. 1990, 24, 1864–1869. [Google Scholar] [CrossRef]
  130. Vempati, H.; Vaitilingom, M.; Zhang, Z.H.; Liyana-Arachchi, T.P.; Stevens, C.S.; Hung, F.R.; Valsararj, K.T. Physico-chemical properties of green leaf volatiles (GLV) for ascertaining atmospheric fate and transport in fog. Adv. Environ. Res. 2018, 7, 139–159. [Google Scholar] [CrossRef]
  131. Buttery, R.G.; Bomben, J.L.; Guadagni, D.G.; Ling, L.C. Volatilities of organic flavor compounds in foods. J. Agric. Food Chem. 1971, 19, 1045–1048. [Google Scholar] [CrossRef]
  132. Nozière, B.; Voisin, D.; Longfellow, C.A.; Friedli, H.; Henry, B.E.; Hanson, D.R. The Uptake of Methyl Vinyl Ketone, Methacrolein, and 2-Methyl-3-butene-2-ol onto Sulfuric Acid Solutions. J. Phys. Chem. A 2006, 110, 2387–2395. [Google Scholar] [CrossRef] [Green Version]
  133. Iraci, L.T.; Baker, B.M.; Tyndall, G.S.; Orlando, J.J. Measurements of the Henry’s Law Coefficients of 2-Methyl-3-buten-2-ol, Methacrolein, and Methylvinyl Ketone. J. Atmos. Chem. 1999, 33, 321–330. [Google Scholar] [CrossRef]
  134. Roberts, D.D.; Pollien, P. Analysis of aroma release during microwave heating. J. Agric. Food Chem. 1997, 45, 4388–4392. [Google Scholar] [CrossRef]
  135. Meylan, W.M.; Howard, P.H. Estimating octanol-air partition coefficients with octanol-water partition coefficients and Henry’s law constants. Chemosphere 2005, 61, 640–644. [Google Scholar] [CrossRef]
  136. Yalkowsky, S.H.; Valvani, S.C. Solubility and partitioning. 1. Solubility of non-electrolytes in water. J. Pharm. Sci. 1980, 69, 912–922. [Google Scholar] [CrossRef] [PubMed]
  137. Riddick, J.A.; Bunger, W.B.; Sakano, T.K. Organic Solvents: Physical Properties and Methods of Purification, 4th ed.; John Wiley and Sons: New York, NY, USA, 1986. [Google Scholar]
  138. Mackay, D.; Yeun, A.T.K. Mass-transfer coefficient correlations for volatilization of organic solutes from water. Environ. Sci. Technol. 1983, 17, 211–217. [Google Scholar] [CrossRef]
  139. Daubert, T.E.; Danner, R.P. Physical and Thermodynamic Properties of Pure Chemicals: Data Compilation; Taylor & Francis: Washington, DC, USA, 1989. [Google Scholar]
  140. Amoore, J.E.; Buttery, R.G. Partition-coefficients and comparative olfactometry. Chem. Sens. Flavour 1978, 3, 57–71. [Google Scholar] [CrossRef]
  141. Yalkowsky, S.H.; Dannenfelser, R. Aquasol Database of Aqueous Solubility, Version 5; College of Pharmacy, University of Arizona: Tuscon, AR, USA, 1992. [Google Scholar]
  142. Yaws, C.L. Handbook of Vapor Pressure; Gulf Pub. Co.: Houston, TX, USA, 1994. [Google Scholar]
  143. Butler, J.A.V.; Ramchandani, C.N.; Thomson, D.W. 58. The solubility of non-electrolytes. Part I. The free energy of hydration of some aliphatic alcohols. J. Chem. Soc. 1935, 280–285. [Google Scholar] [CrossRef]
  144. Buttery, R.G.; Ling, L.; Guadagni, D.G. Food volatiles. Volatilities of aldehydes, ketones, and esters in dilute water solution. J. Agric. Food Chem. 1969, 17, 385–389. [Google Scholar] [CrossRef]
  145. Hwang, Y.L.; Olson, J.D.; Keller, G.E. Steam stripping for removal of organic pollutants from water. 2. Vapor-liquid-equilibrium data. Ind. Eng. Chem. Res. 1992, 31, 1759–1768. [Google Scholar] [CrossRef]
  146. Li, J.J.; Carr, P.W. Measurement of water-hexadecane partition-coefficients by headspace gas-chromatography and calculation of limiting activity-coefficients in water. Anal. Chem. 1993, 65, 1443–1450. [Google Scholar] [CrossRef]
  147. Mackay, D.; Shiu, W.-Y.; Shiu, W.-Y.; Lee, S.C. Handbook of Physical-Chemical Properties and Environmental Fate for Organic Chemicals, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar] [CrossRef]
  148. Rytting, J.H.; Huston, L.P.; Higuchi, T. Thermodynamic group contributions for hydroxyl, amino, and methylene groups. J. Pharm. Sci. 1978, 67, 615–618. [Google Scholar] [CrossRef]
  149. Buttery, R.G.; Guadagni, D.G.; Okano, S. Air-water partition coefficients of some aldehydes. J. Sci. Food Agric. 1965, 16, 691–692. [Google Scholar] [CrossRef]
  150. Sieg, K.; Fries, E.; Puttmann, W. Analysis of benzene, toluene, ethylbenzene, xylenes and n-aldehydes in melted snow water via solid-phase dynamic extraction combined with gas chromatography/mass spectrometry. J. Chromatogr. A 2008, 1178, 178–186. [Google Scholar] [CrossRef] [PubMed]
  151. Gupta, A.K.; Teja, A.S.; Chai, X.S.; Zhu, J.Y. Henry’s constants of n-alkanols (methanol through n-hexanol) in water at temperatures between 40 degrees C and 90 degrees C. Fluid Phase Equilib. 2000, 170, 183–192. [Google Scholar] [CrossRef]
  152. Falabella, J.B.; Nair, A.; Teja, A.S. Henry’s constants of 1-alkanols and 2-ketones in salt solutions. J. Chem. Eng. Data 2006, 51, 1940–1945. [Google Scholar] [CrossRef]
  153. Harvey, A.H. Applications of Near-Critical Dilute-Solution Thermodynamics. Ind. Eng. Chem. Res. 1998, 37, 3080–3088. [Google Scholar] [CrossRef]
  154. Teja, A.S.; Gupta, A.K.; Bullock, K.; Chai, X.-S.; Zhu, J. Henry’s constants of methanol in aqueous systems containing salts. Fluid Phase Equilib. 2001, 185, 265–274. [Google Scholar] [CrossRef]
  155. Dohnal, V.; Fenclova, D.; Vrbka, P. Temperature dependences of limiting activity coefficients, Henry’s law constants, and derivative infinite dilution properties of lower (C-1-C-5) 1-alkanols in water. Critical compilation, correlation, and recommended data. J. Phys. Chem. Ref. Data 2006, 35, 1621–1651. [Google Scholar] [CrossRef]
  156. Liyana-Arachchi, T.P.; Hansel, A.K.; Stevens, C.; Ehrenhauser, F.S.; Valsaraj, K.T.; Hung, F.R. Molecular modeling of the green leaf volatile methyl salicylate on atmospheric air/water interfaces. J. Phys. Chem. A 2013, 117, 4436–4443. [Google Scholar] [CrossRef]
  157. Liyana-Arachchi, T.P.; Zhang, Z.; Vempati, H.; Hansel, A.K.; Stevens, C.; Pham, A.T.; Ehrenhauser, F.S.; Valsaraj, K.T.; Hung, F.R. Green leaf volatiles on atmospheric air/water interfaces: A combined experimental and molecular simulation study. J. Chem. Eng. Data 2014, 59, 3025–3035. [Google Scholar] [CrossRef]
  158. Kerr, J.B.; Fioletov, V.E. Surface ultraviolet radiation. Atmos. Ocean. 2008, 46, 159–184. [Google Scholar] [CrossRef] [Green Version]
  159. Jiménez, E.; Lanza, B.; Martínez, E.; Albaladejo, J. Daytime tropospheric loss of hexanal and trans-2-hexenal: OH kinetics and UV photolysis. Atmos. Chem. Phys. 2007, 7, 1565–1574. [Google Scholar] [CrossRef] [Green Version]
  160. Xing, J.-H.; Ono, M.; Kuroda, A.; Obi, K.; Sato, K.; Imamura, T. Kinetic Study of the Daytime Atmospheric Fate of (Z)-3-Hexenal. J. Phys. Chem. A 2012, 116, 8523–8529. [Google Scholar] [CrossRef] [PubMed]
  161. O’Connor, M.P.; Wenger, J.C.; Mellouki, A.; Wirtz, K.; Munoz, A. The atmospheric photolysis of E-2-hexenal, Z-3-hexenal and E,E-2,4-hexadienal. Phys. Chem. Chem. Phys. 2006, 8, 5236–5246. [Google Scholar] [CrossRef] [PubMed]
  162. Kalalian, C.; Samir, B.; Roth, E.; Chakir, A. UV absorption spectra of trans-2-pentenal, trans-2-hexenal and 2-methyl-2-pentenal. Chem. Phys. Lett. 2019, 718, 22–26. [Google Scholar] [CrossRef]
  163. Sarang, K.; Otto, T.; Rudzinski, K.; Grgić, I.; Nestorowicz, K.; Herrmann, H.; Szmigielski, R. Reaction kinetics of green leaf volatiles with sulfate, hydroxyl and nitrate radicals in tropospheric aqueous-phase. Environ. Sci. Technol. 2021. [Google Scholar] [CrossRef]
  164. Canosa-Mas, C.E.; Duffy, J.M.; King, M.D.; Thompson, K.C.; Wayne, R.P. The atmospheric chemistry of methyl salicylate—reactions with atomic chlorine and with ozone. Atmos. Environ. 2002, 36, 2201–2205. [Google Scholar] [CrossRef]
  165. Rudich, Y.; Talukdar, R.; Burkholder, J.B.; Ravishankara, A.R. Reaction of Methylbutenol with the OH Radical: Mechanism and Atmospheric Implications. J. Phys. Chem. 1995, 99, 12188–12194. [Google Scholar] [CrossRef]
  166. Tang, Y.; Zhu, L. Wavelength-Dependent Photolysis of n-Hexanal and n-Heptanal in the 280−330-nm Region. J. Phys. Chem. A 2004, 108, 8307–8316. [Google Scholar] [CrossRef]
  167. Atkinson, R.; Arey, J. Gas-phase tropospheric chemistry of biogenic volatile organic compounds: A review. Atmos. Environ. 2003, 37 Supp. 2, 197–219. [Google Scholar] [CrossRef]
  168. Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103, 4605–4638. [Google Scholar] [CrossRef] [PubMed]
  169. Vereecken, L.; Francisco, J.S. Theoretical studies of atmospheric reaction mechanisms in the troposphere. Chem. Soc. Rev. 2012, 41, 6259–6293. [Google Scholar] [CrossRef] [PubMed]
  170. Atkinson, R. Atmospheric chemistry of VOCs and NOx. Atmos. Environ. 2000, 34, 2063–2101. [Google Scholar] [CrossRef]
  171. Gai, Y.; Lin, X.; Ma, Q.; Hu, C.; Gu, X.; Zhao, W.; Fang, B.; Zhang, W.; Long, B.; Long, Z. Experimental and Theoretical Study of Reactions of OH Radicals with Hexenols: An Evaluation of the Relative Importance of the H-Abstraction Reaction Channel. Environ. Sci. Technol. 2015, 49, 10380–10388. [Google Scholar] [CrossRef]
  172. Du, B.; Zhang, W. Atmospheric reaction of OH radicals with 2-methyl-3-buten-2-ol (MBO): Quantum chemical investigation on the reaction mechanism. Comput. Theor. Chem. 2012, 999, 109–120. [Google Scholar] [CrossRef]
  173. Davis, M.E.; Gilles, M.K.; Ravishankara, A.R.; Burkholder, J.B. Rate coefficients for the reaction of OH with (E)-2-pentenal, (E)-2-hexenal, and (E)-2-heptenal. Phys. Chem. Chem. Phys. 2007, 9, 2240–2248. [Google Scholar] [CrossRef]
  174. Atkinson, R.; Arey, J.; Aschmann, S.M.; Corchnoy, S.B.; Shu, Y. Rate constants for the gas-phase reactions of cis-3-Hexen-1-ol, cis-3-Hexenylacetate, trans-2-Hexenal, and Linalool with OH and NO3 radicals and O3 at 296 ± 2 K, and OH radical formation yields from the O3 reactions. Int. J. Chem. Kinet. 1995, 27, 941–955. [Google Scholar] [CrossRef]
  175. Gibilisco, R.G.; Blanco, M.a.B.; Bejan, I.; Barnes, I.; Wiesen, P.; Teruel, M.A. Atmospheric Sink of (E)-3-Hexen-1-ol, (Z)-3-Hepten-1-ol, and (Z)-3-Octen-1-ol: Rate Coefficients and Mechanisms of the OH-Radical Initiated Degradation. Environ. Sci. Technol. 2015, 49, 7717–7725. [Google Scholar] [CrossRef] [PubMed]
  176. Gibilisco, R.G.; Santiago, A.N.; Teruel, M.A. OH-initiated degradation of a series of hexenols in the troposphere. Rate coefficients at 298 K and 1 atm. Atmos. Environ. 2013, 77, 358–364. [Google Scholar] [CrossRef]
  177. Carrasco, N.; Doussin, J.F.; O’Connor, M.; Wenger, J.C.; Picquet-Varrault, B.; Durand-Jolibois, R.; Carlier, P. Simulation Chamber Studies of the Atmospheric Oxidation of 2-Methyl-3-Buten-2-ol: Reaction with Hydroxyl Radicals and Ozone Under a Variety of Conditions. J. Atmos. Chem. 2007, 56, 33–55. [Google Scholar] [CrossRef]
  178. Alvarado, A.; Tuazon, E.C.; Aschmann, S.M.; Arey, J.; Atkinson, R. Products and mechanisms of the gas-phase reactions of OH radicals and O3 with 2-methyl-3-buten-2-ol. Atmos. Environ. 1999, 33, 2893–2905. [Google Scholar] [CrossRef]
  179. Ferronato, C.; Orlando, J.J.; Tyndall, G.S. Rate and mechanism of the reactions of OH and Cl with 2-methyl-3-buten-2-ol. J. Geophys. Res. Atmos. 1998, 103, 25579–25586. [Google Scholar] [CrossRef]
  180. Reisen, F.; Aschmann, S.M.; Atkinson, R.; Arey, J. Hydroxyaldehyde Products from Hydroxyl Radical Reactions of Z-3-Hexen-1-ol and 2-Methyl-3-buten-2-ol Quantified by SPME and API-MS. Environ. Sci. Technol. 2003, 37, 4664–4671. [Google Scholar] [CrossRef]
  181. Orlando, J.J.; Tyndall, G.S. Laboratory studies of organic peroxy radical chemistry: An overview with emphasis on recent issues of atmospheric significance. Chem. Soc. Rev. 2012, 41, 6294–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Fantechi, G.; Jensen, N.R.; Hjorth, J.; Peeters, J. Mechanistic studies of the atmospheric oxidation of methyl butenol by OH radicals, ozone and NO3 radicals. Atmos. Environ. 1998, 32, 3547–3556. [Google Scholar] [CrossRef]
  183. Aschmann, S.M.; Shu, Y.; Arey, J.; Atkinson, R. Products of the gas-phase reactions of cis-3-hexen-1-ol with OH radicals and O3. Atmos. Environ. 1997, 31, 3551–3560. [Google Scholar] [CrossRef]
  184. Cavalli, F.; Barnes, I.; Becker, K.H. FT-IR kinetic and product study of the OH radical-initiated oxidation of 1-pentanol. Environ. Sci. Technol. 2000, 34, 4111–4116. [Google Scholar] [CrossRef]
  185. Orlando, J.J.; Tyndall, G.S.; Ceazan, N. Rate coefficients and product yields from reaction of OH with 1-penten-3-ol, (Z)-2-penten-1-ol, and allyl alcohol (2-propen-1-ol). J. Phys. Chem. A 2001, 105, 3564–3569. [Google Scholar] [CrossRef]
  186. Gaona-Colmán, E.; Blanco, M.B.; Teruel, M.A. Kinetics and product identification of the reactions of (E)-2-hexenyl acetate and 4-methyl-3-penten-2-one with OH radicals and Cl atoms at 298 K and atmospheric pressure. Atmos. Environ. 2017, 161, 155–166. [Google Scholar] [CrossRef]
  187. Chan, A.W.H.; Galloway, M.M.; Kwan, A.J.; Chhabra, P.S.; Keutsch, F.N.; Wennberg, P.O.; Flagan, R.C.; Seinfeld, J.H. Photooxidation of 2-Methyl-3-Buten-2-ol (MBO) as a Potential Source of Secondary Organic Aerosol. Environ. Sci. Technol. 2009, 43, 4647–4652. [Google Scholar] [CrossRef]
  188. Bowman, J.H.; Barket, D.J.; Shepson, P.B. Atmospheric chemistry of nonanal. Environ. Sci. Technol. 2003, 37, 2218–2225. [Google Scholar] [CrossRef]
  189. Priya, A.M.; Senthilkumar, L. Reaction of OH radical and ozone with methyl salicylate–a DFT study. J. Phys. Org. Chem. 2015, 28, 542–553. [Google Scholar] [CrossRef]
  190. Seif, A.; Domingo, L.R.; Mazarei, E.; Zahedi, E.; Ahmadi, T.S. Atmospheric Oxidation Reactions of Methyl Salicylate as Green Leaf Volatiles by OH Radical: Theoretical Kinetics and Mechanism. Chem. Sel. 2020, 5, 12535–12547. [Google Scholar] [CrossRef]
  191. Noda, J.; Hallquist, M.; Langer, S.; Ljungström, E. Products from the gas-phase reaction of some unsaturated alcohols with nitrate radicals. Phys. Chem. Chem. Phys. 2000, 2, 2555–2564. [Google Scholar] [CrossRef]
  192. Pfrang, C.; Romero, M.T.B.; Cabanas, B.; Canosa-Mas, C.E.; Villanueva, F.; Wayne, R.P. Night-time tropospheric chemistry of the unsaturated alcohols (Z)-pent-2-en-1-ol and pent-1-en-3-ol: Kinetic studies of reactions of NO3 and N2O5 with stress-induced plant emissions. Atmos. Environ. 2007, 41, 1652–1662. [Google Scholar] [CrossRef]
  193. Kerdouci, J.; Picquet-Varrault, B.; Durand-Jolibois, R.; Gaimoz, C.; Doussin, J.F. An Experimental Study of the Gas-Phase Reactions of NO3 Radicals with a Series of Unsaturated Aldehydes: Trans-2-Hexenal, trans-2-Heptenal, and trans-2-Octenal. J. Phys. Chem. A 2012, 116, 10135–10142. [Google Scholar] [CrossRef] [PubMed]
  194. Horie, O.; Moortgat, G.K. Gas-Phase Ozonolysis of Alkenes. Recent Advances in Mechanistic Investigations. Acc. Chem. Res. 1998, 31, 387–396. [Google Scholar] [CrossRef]
  195. Johnson, D.; Marston, G. The gas-phase ozonolysis of unsaturated volatile organic compounds in the troposphere. Chem. Soc. Rev. 2008, 37, 699–716. [Google Scholar] [CrossRef]
  196. Li, J.; Sun, Y.; Cao, H.; Han, D.; He, M. Mechanisms and kinetics of the ozonolysis reaction of cis-3-hexenyl acetate and trans-2-hexenyl acetate in atmosphere: A theoretical study. Struct. Chem. 2014, 25, 71–83. [Google Scholar] [CrossRef]
  197. Li, J.J. Criegee mechanism of ozonolysis. In Name Reactions: A Collection of Detailed Mechanisms and Synthetic Applications; Springer: Berlin/Heidelberg, Germany, 2009; p. 161. [Google Scholar] [CrossRef]
  198. Jain, S.; Zahardis, J.; Petrucci, G.A. Soft Ionization Chemical Analysis of Secondary Organic Aerosol from Green Leaf Volatiles Emitted by Turf Grass. Environ. Sci. Technol. 2014, 48, 4835–4843. [Google Scholar] [CrossRef] [PubMed]
  199. Grosjean, D.; Grosjean, E. Carbonyl products of the ozone-unsaturated alcohol reaction. J. Geophys. Res. Atmos. 1995, 100, 22815–22820. [Google Scholar] [CrossRef]
  200. Grosjean, E.; Grosjean, D. The Gas Phase Reaction of Unsaturated Oxygenates with Ozone: Carbonyl Products and Comparison with the Alkene-Ozone Reaction. J. Atmos. Chem. 1997, 27, 271–289. [Google Scholar] [CrossRef]
  201. O’Dwyer, M.A.; Carey, T.J.; Healy, R.M.; Wenger, J.C.; Picquet-Varrault, B.; Doussin, J.F. The gas-phase ozonolysis of 1-penten-3-ol, (z)-2-penten-1-ol and 1-penten-3-one: Kinetics, products and secondary organic aerosol formation. Z. FPhys. Chem. 2010, 224, 1059–1080. [Google Scholar] [CrossRef]
  202. Uchida, R.; Sato, K.; Imamura, T. Gas-phase Ozone Reactions with Z-3-Hexenal and Z-3-Hexen-1-ol: Formation Yields of OH Radical, Propanal, and Ethane. Chem. Lett. 2015, 44, 457–458. [Google Scholar] [CrossRef] [Green Version]
  203. Grosjean, E.; Grosjean, D.; Seinfeld, J.H. Gas-phase reaction of ozone with trans-2-hexenal, trans-2-hexenyl acetate, ethylvinyl ketone, and 6-methyl-5-hepten-2-one. Int. J. Chem. Kinet. 1996, 28, 373–382. [Google Scholar] [CrossRef]
  204. Fantechi, G.; Jensen, N.R.; Saastad, O.; Hjorth, J.; Peeters, J. Reactions of Cl Atoms with Selected VOCs: Kinetics, Products and Mechanisms. J. Atmos. Chem. 1998, 31, 247–267. [Google Scholar] [CrossRef]
  205. Rodríguez, A.; Rodríguez, D.; Garzón, A.; Soto, A.; Aranda, A.; Notario, A. Kinetics and mechanism of the atmospheric reactions of atomic chlorine with 1-penten-3-ol and (Z)-2-penten-1-ol: An experimental and theoretical study. Phys. Chem. Chem. Phys. 2010, 12, 12245–12258. [Google Scholar] [CrossRef] [PubMed]
  206. Zhang, W.C.; Zhang, D.J. Theoretical investigation of the oxidation pathways of the Cl-initiated reaction of 2-methyl-3-buten-2-ol. Mol. Phys. 2012, 110, 2901–2917. [Google Scholar] [CrossRef]
  207. Takahashi, K.; Xing, J.H.; Hurley, M.D.; Wallington, T.J. Kinetics and Mechanism of Chlorine-Atom-Initiated Oxidation of Allyl Alcohol, 3-Buten-2-ol, and 2-Methyl-3-buten-2-ol. J. Phys. Chem. A 2010, 114, 4224–4231. [Google Scholar] [CrossRef]
  208. Shashikala, K.; Janardanan, D. Degradation mechanism of trans-2-hexenal in the atmosphere. Chem. Phys. Lett. 2020, 759. [Google Scholar] [CrossRef]
  209. Priya, A.M.; Senthilkumar, L. Degradation of methyl salicylate through Cl initiated atmospheric oxidation—A theoretical study. RSC Adv. 2014, 4, 23464–23475. [Google Scholar] [CrossRef]
  210. Lendar, M.; Aissat, A.; Cazaunau, M.; Daele, V.; Mellouki, A. Absolute and relative rate constants for the reactions of OH and Cl with pentanols. Chem. Phys. Lett. 2013, 582, 38–43. [Google Scholar] [CrossRef] [Green Version]
  211. Wallington, T.J.; Kurylo, M.J. The gas-phase reactions of hydroxyl radicals with a series of aliphatic-alcohols over the temperature-range 240-440-K. Int. J. Chem. Kinet. 1987, 19, 1015–1023. [Google Scholar] [CrossRef]
  212. Nelson, L.; Rattigan, O.; Neavyn, R.; Sidebottom, H.; Treacy, J.; Nielsen, O.J. Absolute and relative rate constants for the reactions of hydroxyl radicals and chlorine atoms with a series of aliphatic-alcohols and ethers at 298-K. Int. J. Chem. Kinet. 1990, 22, 1111–1126. [Google Scholar] [CrossRef]
  213. Wu, H.; Mu, Y.J.; Zhang, X.S.; Jiang, G.B. Relative rate constants for the reactions of hydroxyl radicals and chlorine atoms with a series of aliphatic alcohols. Int. J. Chem. Kinet. 2003, 35, 81–87. [Google Scholar] [CrossRef]
  214. Oh, S.; Andino, J.M. Kinetics of the gas-phase reactions of hydroxyl radicals with C-1-C-6 aliphatic alcohols in the presence of ammonium sulfate aerosols. Int. J. Chem. Kinet. 2001, 33, 422–430. [Google Scholar] [CrossRef]
  215. Grosjean, D.; Williams, E.L. Environmental persistence of organic compounds estimated from structure-reactivity and linear free-energy relationships. Unsaturated aliphatics. Atmos. Environment. Part A Gen. Top. 1992, 26, 1395–1405. [Google Scholar] [CrossRef]
  216. Aschmann, S.M.; Atkinson, R. Kinetics of the gas-phase reactions of the OH radical with selected glycol ethers, glycols, and alcohols. Int. J. Chem. Kinet. 1998, 30, 533–540. [Google Scholar] [CrossRef]
  217. Peirone, S.A.; Barrera, J.A.; Taccone, R.A.; Cometto, P.M.; Lane, S.I. Relative rate coefficient measurements of OH radical reactions with (Z)-2-hexen-1-ol and (E)-3-hexen-1-ol under simulated atmospheric conditions. Atmos. Environ. 2014, 85, 92–98. [Google Scholar] [CrossRef]
  218. D’Anna, B.; Andresen, Ø.; Gefen, Z.; Nielsen, C.J. Kinetic study of OH and NO3 radical reactions with 14 aliphatic aldehydes. Phys. Chem. Chem. Phys. 2001, 3, 3057–3063. [Google Scholar] [CrossRef]
  219. Papagni, C.; Arey, J.; Atkinson, R. Rate constants for the gas-phase reactions of a series of C3 C6 aldehydes with OH and NO3 radicals. Int. J. Chem. Kinet. 2000, 32, 79–84. [Google Scholar] [CrossRef]
  220. Plagens, H. Untersuchugen zum atmosphärenchemischen Abbau langkettiger Aldehyde. Ph.D. Thesis, Bergische Univerität, Wuppertal, Germany, 2001. [Google Scholar]
  221. Gao, T.T.; Andino, J.M.; Rivera, C.C.; Marquez, M.F. Rate Constants of the Gas-Phase Reactions of OH Radicals with trans-2-Hexenal, trans-2-Octenal, and trans-2-Nonenal. Int. J. Chem. Kinet. 2009, 41, 483–489. [Google Scholar] [CrossRef]
  222. Rodríguez, D.; Rodríguez, A.; Bravo, I.; Garzón, A.; Aranda, A.; Diaz-de-Mera, Y.; Notario, A. Kinetic study of the gas-phase reactions of hydroxyl radicals and chlorine atoms with cis-3-hexenylformate. Int. J. Environ. Sci. Technol. 2015, 12, 2881–2890. [Google Scholar] [CrossRef] [Green Version]
  223. Papagni, C.; Arey, J.; Atkinson, R. Rate constants for the gas-phase reactions of OH radicals with a series of unsaturated alcohols. Int. J. Chem. Kinet. 2001, 33, 142–147. [Google Scholar] [CrossRef]
  224. Fantechi, G.; Jensen, N.R.; Hjorth, J.; Peeters, J. Determination of the rate constants for the gas-phase reactions of methyl butenol with OH radicals, ozone, NO3 radicals, and Cl atoms. Int. J. Chem. Kinet. 1998, 30, 589–594. [Google Scholar] [CrossRef]
  225. Imamura, T.; Iida, Y.; Obi, K.; Nagatani, I.; Nakagawa, K.; Patroescu-Klotz, I.; Hatakeyama, S. Rate coefficients for the gas-phase reactions of OH radicals with methylbutenols at 298 K. Int. J. Chem. Kinet. 2004, 36, 379–385. [Google Scholar] [CrossRef]
  226. Baasandorj, M.; Stevens, P.S. Experimental and Theoretical Studies of the Kinetics of the Reactions of OH and OD with 2-Methyl-3-buten-2-ol between 300 and 415 K at Low Pressure. J. Phys. Chem. A 2007, 111, 640–649. [Google Scholar] [CrossRef] [PubMed]
  227. Ren, Y.G.; McGillen, M.R.; Daele, V.; Casas, J.; Mellouk, A. The fate of methyl salicylate in the environment and its role as signal in multitrophic interactions. Sci. Total Environ. 2020, 749. [Google Scholar] [CrossRef]
  228. Davis, M.E.; Burkholder, J.B. Rate coefficients for the gas-phase reaction of OH with (Z)-3-hexen-1-ol, 1-penten-3-ol, (E)-2-penten-1-ol, and (E)-2-hexen-1-ol between 243 and 404 K. Atmos. Chem. Phys. 2011, 11, 3347–3358. [Google Scholar] [CrossRef] [Green Version]
  229. Atkinson, R. A structure-activity relationship for the estimation of rate constants for the gas-phase reactions of OH radicals with organic compounds. Int. J. Chem. Kinet. 1987, 19, 799–828. [Google Scholar] [CrossRef]
  230. Zhao, Z.J.; Husainy, S.; Smith, G.D. Kinetics studies of the gas-phase reactions of NO3 radicals with series of 1-alkenes, dienes, cycloalkenes, alkenols, and alkenals. J. Phys. Chem. A 2011, 115, 12161–12172. [Google Scholar] [CrossRef] [PubMed]
  231. Pfrang, C.; Martin, R.; Canosa-Mas, C.; Wayne, R. Gas-phase reactions of NO3 and N2O5 with (Z)-hex-4-en-1-ol, (Z)-hex-3-en-1-ol (’leaf alcohol’), (E)-hex-3-en-1-ol, (Z)-hex-2-en-1-ol and (E)-hex-2-en-1-ol. Phys. Chem. Chem. Phys. 2006, 8, 354–363. [Google Scholar] [CrossRef]
  232. D’Anna, B.; Nielsen, J.C. Kinetic study of the vapour-phase reaction between aliphatic aldehydes and the nitrate radical. J. Chem. Soc. Faraday Trans. 1997, 93, 3479–3483. [Google Scholar] [CrossRef]
  233. Noda, J.; Holm, C.; Nyman, G.; Langer, S.; Ljungstrom, E. Kinetics of the gas-phase reaction of n-C-6-C-10 aldehydes with the nitrate radical. Int. J. Chem. Kinet. 2003, 35, 120–129. [Google Scholar] [CrossRef]
  234. Hallquist, M.; Langer, S.; Ljungström, E.; Wängberg, I. Rates of reaction between the nitrate radical and some unsaturated alcohols. Int. J. Chem. Kinet. 1996, 28, 467–474. [Google Scholar] [CrossRef]
  235. Cabañas, B.; Salgado, S.; Martín, P.; Baeza, M.T.; Martínez, E. Night-time Atmospheric Loss Process for Unsaturated Aldehydes: Reaction with NO3 Radicals. J. Phys. Chem. A 2001, 105, 4440–4445. [Google Scholar] [CrossRef]
  236. Rudich, Y.; Talukdar, R.K.; Fox, R.W.; Ravishankara, A.R. Rate Coefficients for Reactions of NO3 with a Few Olefins and Oxygenated Olefins. J. Phys. Chem. 1996, 100, 5374–5381. [Google Scholar] [CrossRef]
  237. Grosjean, E.; Grosjean, D. Rate constants for the gas-phase reactions of ozone with unsaturated aliphatic-alcohols. Int. J. Chem. Kinet. 1994, 26, 1185–1191. [Google Scholar] [CrossRef]
  238. Lin, X.; Ma, Q.; Yang, C.; Tang, X.; Zhao, W.; Hu, C.; Gu, X.; Fang, B.; Gai, Y.; Zhang, W. Kinetics and mechanisms of gas phase reactions of hexenols with ozone. RSC Adv. 2016, 6, 83573–83580. [Google Scholar] [CrossRef]
  239. Gibilisco, R.G.; Bejan, I.; Barnes, I.; Wiesen, P.; Teruel, M.A. FTIR gas kinetic study of the reactions of ozone with a series of hexenols at atmospheric pressure and 298K. Chem. Phys. Lett. 2015, 618, 114–118. [Google Scholar] [CrossRef]
  240. Grosjean, D.; Grosjean, E.; Williams II, E.L. Rate constants for the gas-phase reactions of ozone with unsaturated alcohols, esters, and carbonyls. Int. J. Chem. Kinet. 1993, 25, 783–794. [Google Scholar] [CrossRef]
  241. Zhang, Q.; Lin, X.; Gai, Y.; Ma, Q.; Zhao, W.; Fang, B.; Long, B.; Zhang, W. Kinetic and mechanistic study on gas phase reactions of ozone with a series of cis-3-hexenyl esters. RSC Adv. 2018, 8, 4230–4238. [Google Scholar] [CrossRef] [Green Version]
  242. Klawatsch-Carrasco, N.; Doussin, J.F.; Carlier, P. Absolute rate constants for the gas-phase ozonolysis of isoprene and methylbutenol. Int. J. Chem. Kinet. 2004, 36, 152–156. [Google Scholar] [CrossRef]
  243. Kalalian, C.; El Dib, G.; Singh, H.J.; Rao, P.K.; Roth, E.; Chakir, A. Temperature dependent kinetic study of the gas phase reaction of ozone with 1-penten-3-ol, cis-2-penten-1-ol and trans-3-hexen-1-ol: Experimental and theoretical data. Atmos. Environ. 2020, 223, 117306. [Google Scholar] [CrossRef]
  244. Kalalian, C.; Roth, E.; Chakir, A. Rate Coefficients for the Gas-Phase Reaction of Ozone with C5 and C6 Unsaturated Aldehydes. Int. J. Chem. Kinet. 2018, 50, 47–56. [Google Scholar] [CrossRef]
  245. Grira, A.; Amarandei, C.; Romanias, M.N.; El Dib, G.; Canosa, A.; Arsene, C.; Bejan, I.G.; Olariu, R.I.; Coddeville, P.; Tomas, A. Kinetic Measurements of Cl Atom Reactions with C5–C8 Unsaturated Alcohols. Atmosphere 2020, 11, 256. [Google Scholar] [CrossRef] [Green Version]
  246. Rodriguez, D.; Rodriguez, A.; Notario, A.; Aranda, A.; Diaz-de-Mera, Y.; Martinez, E. Kinetic study of the gas-phase reaction of atomic chlorine with a series of aldehydes. Atmos. Chem. Phys. 2005, 5, 3433–3440. [Google Scholar] [CrossRef] [Green Version]
  247. Teruel, M.A.; Achad, M.; Blanco, M.B. Kinetic study of the reactions of Cl atoms with alpha, beta-unsaturated carbonyl compounds at atmospheric pressure and structure activity relations (SARs). Chem. Phys. Lett. 2009, 479, 25–29. [Google Scholar] [CrossRef]
  248. Kaiser, E.W.; Wallington, T.J.; Hurley, M.D. Products and Mechanism of the Reaction of Chlorine Atoms with 3-Pentanone in 700-950 Torr of N-2/O-2 Diluent at 297-515 K. J. Phys. Chem. A 2010, 114, 343–354. [Google Scholar] [CrossRef]
  249. Gibilisco, R.G.; Bejan, I.; Barnes, I.; Wiesen, P.; Teruel, M.A. Rate coefficients at 298 K and 1 atm for the tropospheric degradation of a series of C6, C7 and C8 biogenic unsaturated alcohols initiated by Cl atoms. Atmos. Environ. 2014, 94, 564–572. [Google Scholar] [CrossRef]
  250. Cuevas, C.A.; Notario, A.; Martinez, E.; Albaladejo, J. Temperature-dependence study of the gas-phase reactions of atmospheric Cl atoms with a series of aliphatic aldehydes. Atmos. Environ. 2006, 40, 3845–3854. [Google Scholar] [CrossRef]
  251. Rodríguez, A.; Rodríguez, D.; Soto, A.; Notario, A.; Aranda, A.; Díaz-de-Mera, Y.; Bravo, I. Relative rate measurements of reactions of unsaturated alcohols with atomic chlorine as a function of temperature. Atmos. Environ. 2007, 41, 4693–4702. [Google Scholar] [CrossRef]
  252. NCAR. Tropospheric Ultra Visible Model. Available online: https://www2.acom.ucar.edu/modeling/tropospheric-ultraviolet-and-visible-tuv-radiation-model (accessed on 8 August 2021).
  253. Kwok, E.S.C.; Atkinson, R. Estimation of hydroxyl radical reaction rate constants for gas-phase organic compounds using a structure-reactivity relationship: An update. Atmos. Environ. 1995, 29, 1685–1695. [Google Scholar] [CrossRef]
  254. Peeters, J.; Boullart, W.; Pultau, V.; Vandenberk, S.; Vereecken, L. Structure−Activity Relationship for the Addition of OH to (Poly)alkenes: Site-Specific and Total Rate Constants. J. Phys. Chem. A 2007, 111, 1618–1631. [Google Scholar] [CrossRef] [PubMed]
  255. King, M.D.; Canosa-Mas, C.E.; Wayne, R.P. A structure–activity relationship (SAR) for predicting rate constants for the reaction of NO3, OH and O3 with monoalkenes and conjugated dienes. Phys. Chem. Chem. Phys. 1999, 1, 2239–2246. [Google Scholar] [CrossRef]
  256. Pfrang, C.; King, M.D.; Canosa-Mas, C.E.; Wayne, R.P. Structure–activity relations (SARs) for gas-phase reactions of NO3, OH and O3 with alkenes: An update. Atmos. Environ. 2006, 40, 1180–1186. [Google Scholar] [CrossRef]
  257. Pfrang, C.; King, M.D.; Canosa-Mas, C.E.; Wayne, R.P. Correlations for gas-phase reactions of NO3, OH and O3 with alkenes: An update. Atmos. Environ. 2006, 40, 1170–1179. [Google Scholar] [CrossRef]
  258. Pfrang, C.; King, M.D.; Braeckevelt, M.; Canosa-Mas, C.E.; Wayne, R.P. Gas-phase rate coefficients for reactions of NO3, OH, O3 and O(3P) with unsaturated alcohols and ethers: Correlations and structure–activity relations (SARs). Atmos. Environ. 2008, 42, 3018–3034. [Google Scholar] [CrossRef]
  259. Kerdouci, J.; Picquet-Varrault, B.; Doussin, J.-F. Prediction of Rate Constants for Gas-Phase Reactions of Nitrate Radical with Organic Compounds: A New Structure–Activity Relationship. Chem. Phys. Chem. 2010, 11, 3909–3920. [Google Scholar] [CrossRef]
  260. Ezell, M.J.; Wang, W.; Ezell, A.A.; Soskin, G.; Finlayson-Pitts, B.J. Kinetics of reactions of chlorine atoms with a series of alkenes at 1 atm and 298 K: Structure and reactivity. Phys. Chem. Chem. Phys. 2002, 4, 5813–5820. [Google Scholar] [CrossRef] [Green Version]
  261. McGillen, M.R.; Archibald, A.T.; Carey, T.; Leather, K.E.; Shallcross, D.E.; Wenger, J.C.; Percival, C.J. Structure–activity relationship (SAR) for the prediction of gas-phase ozonolysis rate coefficients: An extension towards heteroatomic unsaturated species. Phys. Chem. Chem. Phys. 2011, 13, 2842–2849. [Google Scholar] [CrossRef] [PubMed]
  262. Doussin, J.F.; Monod, A. Structure–activity relationship for the estimation of OH-oxidation rate constants of carbonyl compounds in the aqueous phase. Atmos. Chem. Phys. 2013, 13, 11625–11641. [Google Scholar] [CrossRef] [Green Version]
  263. Monod, A.; Doussin, J.F. Structure-activity relationship for the estimation of OH-oxidation rate constants of aliphatic organic compounds in the aqueous phase: Alkanes, alcohols, organic acids and bases. Atmos. Environ. 2008, 42, 7611–7622. [Google Scholar] [CrossRef]
  264. Monod, A.; Poulain, L.; Grubert, S.; Voisin, D.; Wortham, H. Kinetics of OH-initiated oxidation of oxygenated organic compounds in the aqueous phase: New rate constants, structure–activity relationships and atmospheric implications. Atmos. Environ. 2005, 39, 7667–7688. [Google Scholar] [CrossRef]
  265. Richards-Henderson, N.K.; Pham, A.T.; Kirk, B.B.; Anastasio, C. Secondary organic aerosol from aqueous reactions of green leaf volatiles with organic triplet excited states and singlet molecular oxygen. Environ. Sci. Technol. 2015, 49, 268–276. [Google Scholar] [CrossRef] [PubMed]
  266. Mael, L.E.; Jacobs, M.I.; Elrod, M.J. Organosulfate and Nitrate Formation and Reactivity from Epoxides Derived from 2-Methyl-3-buten-2-ol. J. Phys. Chem. A 2015, 119, 4464–4472. [Google Scholar] [CrossRef]
  267. Liu, Z.; Ge, M.; Wang, W.; Yin, S.; Tong, S. The uptake of 2-methyl-3-buten-2-ol into aqueous mixed solutions of sulfuric acid and hydrogen peroxide. Phys. Chem. Chem. Phys. 2011, 13, 2069–2075. [Google Scholar] [CrossRef]
  268. Ren, H.; Sedlak, J.A.; Elrod, M.J. General Mechanism for Sulfate Radical Addition to Olefinic Volatile Organic Compounds in Secondary Organic Aerosol. Environ. Sci. Technol. 2021, 55, 1456–1465. [Google Scholar] [CrossRef]
  269. Liyana-Arachchi, T.P.; Stevens, C.; Hansel, A.K.; Ehrenhauser, F.S.; Valsaraj, K.T.; Hung, F.R. Molecular simulations of green leaf volatiles and atmospheric oxidants on air/water interfaces. Phys. Chem. Chem. Phys. 2013, 15, 3583–3592. [Google Scholar] [CrossRef] [PubMed]
  270. Heath, A.; Valsaraj, K.T. An experimental study of the atmospheric oxidation of a biogenic organic compound (methyl jasmonate) in a thin water film as in fog or aerosols. Open J. Air Pollut. 2017, 6, 44–51. [Google Scholar] [CrossRef] [Green Version]
  271. Hansel, A.K. Reaction of Green Leaf Volatiles as a Source of Secondary Organic Aerosols in Fog Droplets. Master’s Thesis, Louisiana State University, Baton Rouge, LA, USA, 2013. [Google Scholar]
  272. Hamilton, J.F.; Lewis, A.C.; Carey, T.J.; Wenger, J.C. Characterization of polar compounds and oligomers in secondary organic aerosol using liquid chromatography coupled to mass spectrometry. Anal. Chem. 2008, 80, 474–480. [Google Scholar] [CrossRef]
  273. Jaoui, M.; Kleindienst, T.E.; Offenberg, J.H.; Lewandowski, M.; Lonneman, W.A. SOA formation from the atmospheric oxidation of 2-methyl-3-buten-2-ol and its implications for PM2.5. Atmos. Chem. Phys. 2012, 12, 2173–2188. [Google Scholar] [CrossRef] [Green Version]
  274. Novelli, A.; Kaminski, M.; Rolletter, M.; Acir, I.H.; Bohn, B.; Dorn, H.P.; Li, X.; Lutz, A.; Nehr, S.; Rohrer, F.; et al. Evaluation of OH and HO2 concentrations and their budgets during photooxidation of 2-methyl-3-butene-2-ol (MBO) in the atmospheric simulation chamber SAPHIR. Atmos. Chem. Phys. 2018, 18, 11409–11422. [Google Scholar] [CrossRef] [Green Version]
  275. 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, 1084–1103. [Google Scholar] [CrossRef]
  276. Edney, E.O.; Kleindienst, T.E.; Conver, T.S.; McIver, C.D.; Corse, E.W.; Weathers, W.S. Polar organic oxygenates in PM2.5 at a southeastern site in the United States. Atmos. Environ. 2003, 37, 3947–3965. [Google Scholar] [CrossRef]
  277. 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, 3191–3206. [Google Scholar] [CrossRef] [Green Version]
  278. Hansen, A.M.K.; Kristensen, K.; Nguyen, Q.T.; Zare, A.; Cozzi, F.; Nøjgaard, J.K.; Skov, H.; Brandt, J.; Christensen, J.H.; Ström, J.; et al. Organosulfates and organic acids in Arctic aerosols: Speciation, annual variation and concentration levels. Atmos. Chem. Phys. 2014, 14, 7807–7823. [Google Scholar] [CrossRef] [Green Version]
  279. 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, 141–151. [Google Scholar] [CrossRef]
  280. Nguyen, Q.T.; Christensen, M.K.; Cozzi, F.; Zare, A.; Hansen, A.M.K.; Kristensen, K.; Tulinius, T.E.; Madsen, H.H.; Christensen, J.H.; Brandt, J.; et al. Understanding the anthropogenic influence on formation of biogenic secondary organic aerosols in Denmark via analysis of organosulfates and related oxidation products. Atmos. Chem. Phys. 2014, 14, 8961–8981. [Google Scholar] [CrossRef] [Green Version]
  281. Kristensen, K.; Bilde, M.; Aalto, P.P.; Petaja, T.; Glasius, M. Denuder/filter sampling of organic acids and organosulfates at urban and boreal forest sites: Gas/particle distribution and possible sampling artifacts. Atmos. Environ. 2016, 130, 36–53. [Google Scholar] [CrossRef] [Green Version]
  282. 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, 2347–2358. [Google Scholar] [CrossRef] [Green Version]
  283. 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, 1343–1359. [Google Scholar] [CrossRef] [Green Version]
  284. Wang, Y.; Hu, M.; Guo, S.; Wang, Y.; Zheng, J.; Yang, Y.; Zhu, W.; Tang, R.; Li, X.; Liu, Y.; et al. The secondary formation of organosulfates under interactions between biogenic emissions and anthropogenic pollutants in summer in Beijing. Atmos. Chem. Phys. 2018, 18, 10693–10713. [Google Scholar] [CrossRef] [Green Version]
  285. Huang, R.J.; Cao, J.; Chen, Y.; Yang, L.; Shen, J.; You, Q.; Wang, K.; Lin, C.; Xu, W.; Gao, B.; et al. Organosulfates in atmospheric aerosol: Synthesis and quantitative analysis of PM2.5 from Xi’an, northwestern China. Atmos. Meas. Tech. 2018, 11, 3447–3456. [Google Scholar] [CrossRef] [Green Version]
  286. Le Breton, M.; Wang, Y.; Hallquist, Å.M.; Pathak, R.K.; Zheng, J.; Yang, Y.; Shang, D.; Glasius, M.; Bannan, T.J.; Liu, Q.; et al. Online gas- and particle-phase measurements of organosulfates, organosulfonates and nitrooxy organosulfates in Beijing utilizing a FIGAERO ToF-CIMS. Atmos. Chem. Phys. 2018, 18, 10355–10371. [Google Scholar] [CrossRef] [Green Version]
  287. 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, 6468–6474. [Google Scholar] [CrossRef]
  288. Hughes, D.D.; Christiansen, M.B.; Milani, A.; Vermeuel, M.P.; Novak, G.A.; Alwe, H.D.; Dickens, A.F.; Pierce, R.B.; Millet, D.B.; Bertram, T.H.; et al. PM2.5 chemistry, organosulfates, and secondary organic aerosol during the 2017 Lake Michigan Ozone Study. Atmos. Environ. 2021, 244, 117939. [Google Scholar] [CrossRef]
  289. Shalamzari, M.S.; Vermeylen, R.; Blockhuys, F.; Kleindienst, T.E.; Lewandowski, M.; Szmigielski, R.; Rudzinski, 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, 7135–7148. [Google Scholar] [CrossRef] [Green Version]
  290. Safi Shalamzari, M.; Ryabtsova, O.; Kahnt, A.; Vermeylen, R.; Hérent, 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, 784–794. [Google Scholar] [CrossRef] [PubMed]
  291. Brüggemann, M.; Xu, R.; 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, 3767–3782. [Google Scholar] [CrossRef]
  292. Tolocka, M.P.; Turpin, B. Contribution of Organosulfur Compounds to Organic Aerosol Mass. Environ. Sci. Technol. 2012, 46, 7978–7983. [Google Scholar] [CrossRef] [PubMed]
  293. Zhang, H.; 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, 9437–9446. [Google Scholar] [CrossRef] [PubMed]
  294. 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, 12671–12678. [Google Scholar] [CrossRef]
  295. Faiola, C.L.; Pullinen, I.; Buchholz, A.; Khalaj, F.; Ylisirniö, A.; Kari, E.; Miettinen, P.; Holopainen, J.K.; Kivimäenpää, M.; Schobesberger, S.; et al. Secondary Organic Aerosol Formation from Healthy and Aphid-Stressed Scots Pine Emissions. ACS Earth Space Chem. 2019, 3, 1756–1772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Waza, A.A. Effect of Green Leaf Volatiles (GLVS) on Secondary Organic Aerosol Formation; University of Eastern Finland: Joensuu, Finland, 2014. [Google Scholar]
  297. Joutsensaari, J.; Loivamäki, M.; Vuorinen, T.; Miettinen, P.; Nerg, A.M.; Holopainen, J.K.; Laaksonen, A. Nanoparticle formation by ozonolysis of inducible plant volatiles. Atmos. Chem. Phys. 2005, 5, 1489–1495. [Google Scholar] [CrossRef] [Green Version]
  298. Zhang, H.; Zhang, Z.; Cui, T.; Lin, Y.-H.; Bhathela, N.A.; Ortega, J.; Worton, D.R.; Goldstein, A.H.; Guenther, A.; Jimenez, J.L.; et al. Secondary Organic Aerosol Formation via 2-Methyl-3-buten-2-ol Photooxidation: Evidence of Acid-Catalyzed Reactive Uptake of Epoxides. Environ. Sci. Technol. Lett. 2014, 1, 242–247. [Google Scholar] [CrossRef] [PubMed]
  299. Knap, H.C.; Schmidt, J.A.; Jørgensen, S. Hydrogen shift reactions in four methyl-buten-ol (MBO) peroxy radicals and their impact on the atmosphere. Atmos. Environ. 2016, 147, 79–87. [Google Scholar] [CrossRef]
  300. Gomez-Gonzalez, 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, 371–382. [Google Scholar] [CrossRef]
  301. 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, 2129–2140. [Google Scholar] [CrossRef] [Green Version]
  302. Wach, P.; Spolnik, 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, 1415–1424. [Google Scholar] [CrossRef]
  303. 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, 1–9. [Google Scholar] [CrossRef]
  304. Rudzinski, K. Heterogeneous and aqueous-phase transformations of isoprene. In Environmental Simulation Chambers: Application to Atmospheric Chemical Processes; Barnes, I., Rudzinski, K., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; Volume 62, pp. 261–277. [Google Scholar]
  305. Went, F.W. Blue Hazes in the Atmosphere. Nature 1960, 187, 641–643. [Google Scholar] [CrossRef]
  306. Herrmann, H.; Hoffmann, D.; Schaefer, T.; Bräuer, P.; Tilgner, A. Tropospheric aqueous-phase free-radical chemistry: Radical sources, spectra, reaction kinetics and prediction tools. Chem. Phys. Chem. 2010, 11, 3796–3822. [Google Scholar] [CrossRef] [PubMed]
  307. White, R.P.; Murray, S.; Rohweder, M. Grassland Ecosystems; World Research Institute: Washington, DC, USA, 2000. [Google Scholar]
  308. Kang, S.; Nair, S.S.; Kline, K.L.; Nichols, J.A.; Wang, D.; Post, W.M.; Brandt, C.C.; Wullschleger, S.D.; Singh, N.; Wei, Y. Global simulation of bioenergy crop productivity: Analytical framework and case study for switchgrass. GCB Bioenergy 2014, 6, 14–25. [Google Scholar] [CrossRef] [Green Version]
  309. Fan, P.; Hao, M.; Ding, F.; Jiang, D.; Dong, D. Quantifying Global Potential Marginal Land Resources for Switchgrass. Energies 2020, 13, 6197. [Google Scholar] [CrossRef]
Scheme 1. Conversion of linoleic acid into hexenals, hexenols, and hexenyl acetates [84].
Scheme 1. Conversion of linoleic acid into hexenals, hexenols, and hexenyl acetates [84].
Atmosphere 12 01655 sch001
Scheme 2. Formation pathway of C5 GLV [9].
Scheme 2. Formation pathway of C5 GLV [9].
Atmosphere 12 01655 sch002
Scheme 3. Formation pathway of jasmonic acid [32].
Scheme 3. Formation pathway of jasmonic acid [32].
Atmosphere 12 01655 sch003
Scheme 4. Reactions that most often occur when alkene reacts with an atmospheric oxidant X (OH, NO3, O3, Cl) in the presence of oxygen and nitrogen oxides [168].
Scheme 4. Reactions that most often occur when alkene reacts with an atmospheric oxidant X (OH, NO3, O3, Cl) in the presence of oxygen and nitrogen oxides [168].
Atmosphere 12 01655 sch004
Scheme 5. Hydrogen abstraction from a GLV aldehyde and formation of a peroxy radical [173].
Scheme 5. Hydrogen abstraction from a GLV aldehyde and formation of a peroxy radical [173].
Atmosphere 12 01655 sch005
Scheme 6. Reactions of the peroxy radicals with NO2 and NO produce a peroxynitrate and an alkoxy radical, resp. [173].
Scheme 6. Reactions of the peroxy radicals with NO2 and NO produce a peroxynitrate and an alkoxy radical, resp. [173].
Atmosphere 12 01655 sch006
Scheme 7. Reactions of the alkoxy radical leading to the formation of an aldehyde and an alkylglyoxal [173].
Scheme 7. Reactions of the alkoxy radical leading to the formation of an aldehyde and an alkylglyoxal [173].
Atmosphere 12 01655 sch007
Scheme 8. The OH radical addition to the C=C bond in (Z)-3-hexen-1-ol [95].
Scheme 8. The OH radical addition to the C=C bond in (Z)-3-hexen-1-ol [95].
Atmosphere 12 01655 sch008
Scheme 9. Self-reaction of peroxy radicals [95] and other references in this section.
Scheme 9. Self-reaction of peroxy radicals [95] and other references in this section.
Atmosphere 12 01655 sch009
Scheme 10. Reactions of peroxy radicals from (Z)-3-hexen-1-ol with NO lead to the formation of organic nitrates and alkoxy radicals [95,173,174,175,176,177,178,179,180,181,182].
Scheme 10. Reactions of peroxy radicals from (Z)-3-hexen-1-ol with NO lead to the formation of organic nitrates and alkoxy radicals [95,173,174,175,176,177,178,179,180,181,182].
Atmosphere 12 01655 sch010
Scheme 11. Reactions of alkoxy radicals derived from (Z)-3-hexen-1-ol [95,174,175,176,180,183].
Scheme 11. Reactions of alkoxy radicals derived from (Z)-3-hexen-1-ol [95,174,175,176,180,183].
Atmosphere 12 01655 sch011
Scheme 12. Reactions of MBO-derived alkoxy radicals: 2,3-decomposition, 3,4-decompositions, and isomerization [177,178,180,182].
Scheme 12. Reactions of MBO-derived alkoxy radicals: 2,3-decomposition, 3,4-decompositions, and isomerization [177,178,180,182].
Atmosphere 12 01655 sch012
Scheme 13. Reactions of MBO-derived peroxy radicals with NO produce MW 165 dihydroxynitrates [180].
Scheme 13. Reactions of MBO-derived peroxy radicals with NO produce MW 165 dihydroxynitrates [180].
Atmosphere 12 01655 sch013
Scheme 14. The mechanism of OH radical addition to GLV aldehydes includes several decompositions of alkoxy radicals [173].
Scheme 14. The mechanism of OH radical addition to GLV aldehydes includes several decompositions of alkoxy radicals [173].
Atmosphere 12 01655 sch014
Scheme 15. Dominating hydrogen abstraction channel in the reaction of MeSa with OH radicals [189].
Scheme 15. Dominating hydrogen abstraction channel in the reaction of MeSa with OH radicals [189].
Atmosphere 12 01655 sch015
Scheme 16. Dominating addition channel in the reaction of MeSa with OH radicals [189].
Scheme 16. Dominating addition channel in the reaction of MeSa with OH radicals [189].
Atmosphere 12 01655 sch016
Scheme 17. Possible pathways in reactions of GLV with NO3 radicals (for simplicity, the scheme shows the transformation of only one isomeric alkyl radical) [168,170].
Scheme 17. Possible pathways in reactions of GLV with NO3 radicals (for simplicity, the scheme shows the transformation of only one isomeric alkyl radical) [168,170].
Atmosphere 12 01655 sch017
Scheme 18. Mechanism of MBO reaction with NO3 radicals (for simplicity, the scheme shows the transformation of only one isomeric alkyl radical) [182].
Scheme 18. Mechanism of MBO reaction with NO3 radicals (for simplicity, the scheme shows the transformation of only one isomeric alkyl radical) [182].
Atmosphere 12 01655 sch018
Scheme 19. Alternative decomposition steps of MBO-derived nitrate alkoxy radicals [191].
Scheme 19. Alternative decomposition steps of MBO-derived nitrate alkoxy radicals [191].
Atmosphere 12 01655 sch019
Scheme 20. General mechanism of the O3 reactions with alkenes, including GLV [195,197].
Scheme 20. General mechanism of the O3 reactions with alkenes, including GLV [195,197].
Atmosphere 12 01655 sch020
Scheme 21. The basic mechanism of the O3 reaction with (Z)-3-hexenol [92,198].
Scheme 21. The basic mechanism of the O3 reaction with (Z)-3-hexenol [92,198].
Atmosphere 12 01655 sch021
Scheme 22. Proposed hydroperoxide channels in the O3 reaction with (Z)-3-hexenol [198].
Scheme 22. Proposed hydroperoxide channels in the O3 reaction with (Z)-3-hexenol [198].
Atmosphere 12 01655 sch022
Scheme 23. The basic mechanism of the O3 reaction with (Z)-3-hexenyl acetate [92,196,198,199,200].
Scheme 23. The basic mechanism of the O3 reaction with (Z)-3-hexenyl acetate [92,196,198,199,200].
Atmosphere 12 01655 sch023
Scheme 24. Theoretically predicted stabilization channels of Criegee intermediates from (Z)-3-hexenyl acetate [196].
Scheme 24. Theoretically predicted stabilization channels of Criegee intermediates from (Z)-3-hexenyl acetate [196].
Atmosphere 12 01655 sch024
Scheme 25. More theoretically predicted stabilization paths of Criegee intermediates from (Z)-3-hexenyl acetate [196].
Scheme 25. More theoretically predicted stabilization paths of Criegee intermediates from (Z)-3-hexenyl acetate [196].
Atmosphere 12 01655 sch025
Scheme 26. Mechanism of MBO reaction with ozone [177,182].
Scheme 26. Mechanism of MBO reaction with ozone [177,182].
Atmosphere 12 01655 sch026
Scheme 27. Simplified mechanism of MeSa ozonolysis [189].
Scheme 27. Simplified mechanism of MeSa ozonolysis [189].
Atmosphere 12 01655 sch027
Scheme 28. The Cl radical addition to the C=C bond in 1-penten-3-ol followed by isomerization or reaction with molecular oxygen [205].
Scheme 28. The Cl radical addition to the C=C bond in 1-penten-3-ol followed by isomerization or reaction with molecular oxygen [205].
Atmosphere 12 01655 sch028
Scheme 29. The self-reaction and isomerization peroxy radicals derived from 1-penten-3-ol in the absence of NO [205].
Scheme 29. The self-reaction and isomerization peroxy radicals derived from 1-penten-3-ol in the absence of NO [205].
Atmosphere 12 01655 sch029
Scheme 30. Reactions of alkoxy radicals in the mechanism of the reaction of 1-penten-3-ol with Cl. [205].
Scheme 30. Reactions of alkoxy radicals in the mechanism of the reaction of 1-penten-3-ol with Cl. [205].
Atmosphere 12 01655 sch030
Scheme 31. Hydrogen abstraction from 1-penten-3-ol by a Cl radical, followed by the formation of a peroxy radical and a self-reaction leading to the formation of 1-penten-3-one [205].
Scheme 31. Hydrogen abstraction from 1-penten-3-ol by a Cl radical, followed by the formation of a peroxy radical and a self-reaction leading to the formation of 1-penten-3-one [205].
Atmosphere 12 01655 sch031
Scheme 32. Reaction of alkoxy radicals produced in the addition and hydrogen abstraction channel of the (Z)-2-penten-1-ol reaction with Cl radicals [205].
Scheme 32. Reaction of alkoxy radicals produced in the addition and hydrogen abstraction channel of the (Z)-2-penten-1-ol reaction with Cl radicals [205].
Atmosphere 12 01655 sch032
Scheme 33. The reaction of the alkoxy radicals derived from the MBO reactions with Cl (the addition and hydrogen abstraction channels) in the absence of NO [179,206,207].
Scheme 33. The reaction of the alkoxy radicals derived from the MBO reactions with Cl (the addition and hydrogen abstraction channels) in the absence of NO [179,206,207].
Atmosphere 12 01655 sch033
Scheme 34. Reactions of peroxy radicals from the MBO-Cl reaction with NO lead to alkoxy radicals’ formation [204].
Scheme 34. Reactions of peroxy radicals from the MBO-Cl reaction with NO lead to alkoxy radicals’ formation [204].
Atmosphere 12 01655 sch034
Scheme 35. Reactions of alkoxy radicals resulting from the Cl addition to MBO at the C4 position in the presence of NO and O2 [204].
Scheme 35. Reactions of alkoxy radicals resulting from the Cl addition to MBO at the C4 position in the presence of NO and O2 [204].
Atmosphere 12 01655 sch035
Scheme 36. Reactions of alkoxy radicals resulting from the Cl addition to MBO at the C3 position in the presence of NO and O2 [204].
Scheme 36. Reactions of alkoxy radicals resulting from the Cl addition to MBO at the C3 position in the presence of NO and O2 [204].
Atmosphere 12 01655 sch036
Scheme 37. The reaction of an alkoxy radical generated by the hydrogen abstraction from MBO at the methyl group in the presence of NO and O2 [204].
Scheme 37. The reaction of an alkoxy radical generated by the hydrogen abstraction from MBO at the methyl group in the presence of NO and O2 [204].
Atmosphere 12 01655 sch037
Scheme 38. Reaction of alkoxy radicals derived from the (E)-2-hexenyl acetate reaction with Cl radicals: decomposition and addition of molecular oxygen followed by removal of HO2 [186].
Scheme 38. Reaction of alkoxy radicals derived from the (E)-2-hexenyl acetate reaction with Cl radicals: decomposition and addition of molecular oxygen followed by removal of HO2 [186].
Atmosphere 12 01655 sch038
Scheme 39. Dominating hydrogen abstraction channel at the meta position of MeSa and formation of peroxy radicals [209].
Scheme 39. Dominating hydrogen abstraction channel at the meta position of MeSa and formation of peroxy radicals [209].
Atmosphere 12 01655 sch039
Scheme 40. Possible reaction pathways of MeSa-based peroxy radicals in the presence of NO [209].
Scheme 40. Possible reaction pathways of MeSa-based peroxy radicals in the presence of NO [209].
Atmosphere 12 01655 sch040
Scheme 41. Photolysis of (Z)-3-hexenal in the presence of oxygen [161].
Scheme 41. Photolysis of (Z)-3-hexenal in the presence of oxygen [161].
Atmosphere 12 01655 sch041
Scheme 42. Thermodynamically feasible dissociation pathways after UV excitation of n-hexanal [166].
Scheme 42. Thermodynamically feasible dissociation pathways after UV excitation of n-hexanal [166].
Atmosphere 12 01655 sch042
Scheme 43. Formation of 2-methyl-3-buten-2-ol epoxides in the gas phase.
Scheme 43. Formation of 2-methyl-3-buten-2-ol epoxides in the gas phase.
Atmosphere 12 01655 sch043
Scheme 44. Hydrolysis, sulfation, and nitration of MBO 3,4-epoxide [266,267].
Scheme 44. Hydrolysis, sulfation, and nitration of MBO 3,4-epoxide [266,267].
Atmosphere 12 01655 sch044
Scheme 45. Aqueous-phase reactions of hydroperoxides and epoxides from the gas-phase oxidation of (Z)-3-hexenol [279].
Scheme 45. Aqueous-phase reactions of hydroperoxides and epoxides from the gas-phase oxidation of (Z)-3-hexenol [279].
Atmosphere 12 01655 sch045
Scheme 46. Mechanism of SO4 addition to MBO (based on [268]) shows pathways starting from one of two possible isomers of the alkyl sulfate radical.
Scheme 46. Mechanism of SO4 addition to MBO (based on [268]) shows pathways starting from one of two possible isomers of the alkyl sulfate radical.
Atmosphere 12 01655 sch046
Scheme 47. A plausible mechanism of MeJa reactions initiated by OH radicals in the aqueous phase (elemental formulas annotate products identified in laboratory experiments) [21,271].
Scheme 47. A plausible mechanism of MeJa reactions initiated by OH radicals in the aqueous phase (elemental formulas annotate products identified in laboratory experiments) [21,271].
Atmosphere 12 01655 sch047
Scheme 48. A plausible mechanism of MeSa reactions initiated by OH radicals in the aqueous phase (elemental formulas mark the products identified in laboratory experiments) [21,271]. The C8H8O5 products can also form by an alternative path (Scheme 49).
Scheme 48. A plausible mechanism of MeSa reactions initiated by OH radicals in the aqueous phase (elemental formulas mark the products identified in laboratory experiments) [21,271]. The C8H8O5 products can also form by an alternative path (Scheme 49).
Atmosphere 12 01655 sch048
Scheme 49. An alternative mechanism of C8H8O5 formation in the oxidation of MeSa initiated by OH radicals [21,271].
Scheme 49. An alternative mechanism of C8H8O5 formation in the oxidation of MeSa initiated by OH radicals [21,271].
Atmosphere 12 01655 sch049
Scheme 50. Pinacol rearrangement in aqueous transformation of MBO [132].
Scheme 50. Pinacol rearrangement in aqueous transformation of MBO [132].
Atmosphere 12 01655 sch050
Scheme 51. Particle-phase formation of HMW compounds from 3-hydroxypropanal produced by ozonolysis of (Z)-3-hexen-1-ol [65,198].
Scheme 51. Particle-phase formation of HMW compounds from 3-hydroxypropanal produced by ozonolysis of (Z)-3-hexen-1-ol [65,198].
Atmosphere 12 01655 sch051
Scheme 52. Particle-phase oligomerization of 3-oxopropyl acetate and propanal that are produced by the gas-phase ozonolysis of (Z)-3-hexenyl acetate [92].
Scheme 52. Particle-phase oligomerization of 3-oxopropyl acetate and propanal that are produced by the gas-phase ozonolysis of (Z)-3-hexenyl acetate [92].
Atmosphere 12 01655 sch052
Scheme 53. Mechanism of multiphase ozonolysis of 1-octen-3-ol in a smog chamber [48]. Products with specified MW values were identified in chamber SOA samples using NIR-LDI-AMS approach.
Scheme 53. Mechanism of multiphase ozonolysis of 1-octen-3-ol in a smog chamber [48]. Products with specified MW values were identified in chamber SOA samples using NIR-LDI-AMS approach.
Atmosphere 12 01655 sch053
Table 1. Key Green Leaf Volatiles detected in plant emissions a.
Table 1. Key Green Leaf Volatiles detected in plant emissions a.
NameChemical
Formula
MW g mol−1StructureReferences
pentan-1-olC5H12O88.15 Atmosphere 12 01655 i001[9,26]
1-penten-3-olC5H10O86.13 Atmosphere 12 01655 i002[9,16,18,26,27,28,29,30]
(Z)-2-penten-1-olC5H10O86.13 Atmosphere 12 01655 i003[26,29,30,31]
(E)-2-pentenolC5H10O86.13 Atmosphere 12 01655 i004[28]
(E)-2-pentenalC5H8O84.12 Atmosphere 12 01655 i005[28,30]
1-penten-3-oneC5H8O84.12 Atmosphere 12 01655 i006[9,16,27,28,29,30]
(Z)-2-pentenyl acetateC7H12O2128.17 Atmosphere 12 01655 i007[31]
n-hexan-1-olC6H14O102.16 Atmosphere 12 01655 i008[16,26,28,32,33]
(E)-2-hexen-1-olC6H12O100.16 Atmosphere 12 01655 i009[26,28,32]
(Z)-2-hexen-1-olC6H12O100.16 Atmosphere 12 01655 i010[18]
(E)-3-hexen-1-olC6H12O100.16 Atmosphere 12 01655 i011[28,32]
(Z)-3-hexen-1-ol (leaf alcohol)C6H12O100.16 Atmosphere 12 01655 i012[9,10,12,13,16,18,26,28,31,32,34,35,36,37]
n-hexan-1-alC6H12O100.16 Atmosphere 12 01655 i013[3,18,26,28,30,32,33,34,36,38]
(Z)-3-hexen-1-alC6H10O98.14 Atmosphere 12 01655 i014[9,12,13,18,26,28,32,33,34,37,38,39]
(E)-2-hexen-1-al (leaf aldehyde)C6H10O98.14 Atmosphere 12 01655 i015[9,10,18,26,28,31,32,33,35,36,38]
(Z)-2-hexenalC6H10O98.14 Atmosphere 12 01655 i016[16,28]
(E)-3-hexenalC6H10O98.14 Atmosphere 12 01655 i017[28,32,37]
hexyl acetateC8H16O2144.21 Atmosphere 12 01655 i018[33,37,40,41]
(Z)-2-hexenyl acetateC8H14O2142.20 Atmosphere 12 01655 i019[18]
(E)-2-hexenyl acetateC8H14O2142.20 Atmosphere 12 01655 i020[28,37]
(Z)-3-hexenyl acetate (leaf acetate)C8H14O2142.20 Atmosphere 12 01655 i021[12,13,16,18,26,28,31,32,34,37]
(Z)-3-hexenyl-propionateC9H16O2156.22 Atmosphere 12 01655 i022[28,30,42]
(Z)-3-hexenyl butanoate C10H18O2170.25 Atmosphere 12 01655 i023[30,37,40,41,42,43]
(Z)-3-hexenyl isobutanoate (isobutyrate)C10H18O2170.25 Atmosphere 12 01655 i024[30,40,42]
(E)-3-hexenyl butanoateC10H18O2170.25 Atmosphere 12 01655 i025[41]
(E)-2-hexenyl butanoateC10H18O2170.25 Atmosphere 12 01655 i026[37,42]
(Z)-3-hexenyl isopentanateC11H20O2184.27 Atmosphere 12 01655 i027[40,42,44]
(Z)-3-hexenyl 2-methyl-2-butenoateC11H18O2182.26 Atmosphere 12 01655 i028[40,42,44]
3-hexenyl hexanoate bC12H22O2198.3 Atmosphere 12 01655 i029[41]
2-methy-3-buten-2-olC5H10O86.13 Atmosphere 12 01655 i030[45,46,47]
1-octen-3-olC8H16O128.21 Atmosphere 12 01655 i031[18,28,48]
nonanalC9H18O142.24 Atmosphere 12 01655 i032[44]
jasmonic acidC12H18O3210.27 Atmosphere 12 01655 i033[26,39,49]
methyl jasmonateC13H20O3224.30 Atmosphere 12 01655 i034[26,49]
methyl salicylateC8H8O3152.15 Atmosphere 12 01655 i035[7,30,40]
a for convenience of the audience, we used traditional GLV names rather than the latest IUPAC recommendations; b unspecified isomer.
Table 2. Emission and emission rates of atmospherically relevant GLV.
Table 2. Emission and emission rates of atmospherically relevant GLV.
GLVEmission or Emission Rate
pentan-1-ol
  • 0.2 and 8.8 ng g DW−1 h−1 rape and blossoming rye, resp. (field enclosure) [86] a;
    7.15 ± 0.46 pmol m−2 s−1 L. pulmonaria thalii: (intact);
    18.4 ± 0.6, 28.0 ± 3.7, 56.3 ± 7.8 pmol m−2 s−1 after heat shock of 37, 46, and 51 °C, resp.; 68.6 ± 13.9 pmol m−2 s−1 (wounded) [87].
1-penten-3-ol
  • 0.2 kg C ha−1 yr−1 switchgrass plantations (growth and harvest);
    18 ± 9 ng g DW−1 h−1 (6 ± 3 pmol m−2 s−1) mean before cutting;
    91 ± 114 ng g DW−1 h−1 (31 ± 39 pmol m−2 s−1) max after cutting;
    48 ± 22 ng g DW−1 h−1 (17 ± 17 pmol m−2 s−1) mean after cutting (including isoprene) b [88].
  • 0.2, 0.3, 1.2 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16];
  • 17.4, 22.0, 4.8, 5.9, 1.9, and 37.7 ng g DW−1 h−1 rape, blossoming rape, rye, blossoming rye, grassland, and grassland after mowing, resp. (in a field enclosure) [86].
(E)-2-penten-1-ol
  • 0.2 and 1.9 ng g DW−1 h−1 rape and blossoming rye (in a field enclosure) [86].
3-pentanone
  • 0.2, 0.5, 2 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16].
1-hexanol
  • (3.5 ± 4.0), (18.2 ± 4.9), (0.5 ± 1.1) ng g DM−1 Betula pendula, B. pubescens, Populus tremula, resp. (dry leaves, June) [41].
  • < 0.3 and 2.3 pmol m−2 s−1 Bel-W3 tobacco, before and after exposure to O3 in a lab reactor [89].
  • 0.1, 0.2, 0.3 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16].
  • 0.27 ± 0.04 pmol m−2 s−1 L. pulmonaria thalli: (intact);
    0.27 ± 0.6, 0.33 ± 0.03, 0.52 ± 0.06 pmol m−2 s−1 after heat shock of 37, 46, and 51 °C, resp.;
    0.46 ± 0.05 pmol m−2 s−1 (wounded) [87].
  • 58.2 and 31.1 ng g DW−1 h−1 rye and blossoming rye (in a field enclosure) [86].
  • 1.13 ± 1.13 ng g DW−1 h−1 Betula pendula [44].
  • 0.6, 0.3 μg C g DW−1 h−1 rabbit brush, Metasequoia glyptostroboides (cut branches) [90].
  • (0.4 ± 0.4)–(0.5 ± 0.5) ng g DW−1 h−1 Pinus sylvestris (various seedlings) [91].
(E)-2-hexen-1-ol
  • (4.6 ± 1.4 | 1.3 ± 1.0), (4.2 ± 1.8 | 1.3 ± 1.0), (1.5 ± 1.0 | 0.6 ± 0.5), (- |1.0 ± 0.8) ng g DW−1, (June | August) Betula pendula, B. pubescens, Populus tremula, and Sambucus nigra, resp. (dry leaves) [41].
  • 0.1–1.2 μg C g DW−1 h−1 various trees (cut branches) [90].
(Z)-3-hexen-1-ol
  • 1.21 ± 0.06 μg m−2 mowed lawn [92].
  • 9.6 and 163 pmol m−2 s−1 Bel-W3 tobacco, before and after exposure to O3 in a lab reactor [89].
  • 0.10 ± 0.03, 0.039 ± 0.006 pmol m−2 s−1 intact apple and grape foliage day mean, resp. [93].
  • (108 ± 53), (2736 ± 546) ng/30 min from caterpillar-infested corn plant and its cut leaves, resp. (in a chamber) [34].
  • 0.1, 0.3, 0.7 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16].
  • (3.4 ± 6.7), (30 ± 46) μg m−2 h−1 (1992), (1993), resp., from a test grassland in Argonne June–August average at variable locations, noon, sunny days [94].
  • (0.85 ± 2.2), (54 ± 110) μg m−2 h−1 (1992), (1993), resp., from a test grassland in Argonne June-August average at fixed location, noon, sunny days [94].
  • (29.0, 2.4), 20.1, 20.3, 33.9, 11.8, 0.5, 31.5, 48.4, 6.8, (0.8, 6.1, 7.5) and 23.6 ng g DW−1 h−1 (grape with fruits), rape, blossoming rape, rye, blossoming rye, beech, hornbeam, birch, oak, (grassland), grassland after mowing, resp. (in a field enclosure) [86].
  • 0.2, 0.2, 0.06, 0.03, 0.05, 0.2, 0.07, 0.1, 0.06, 0.8, 1.3, 0.3, 0.3, 0.3, and 0.5 μg g DW−1 h−1 alfalfa, almond, apricot, bean, cotton, grape, nectarine, olive, orange, peach, pistachio, plum, walnut, valley oak, and whitehorn, resp. (in flow-through field enclosures) [95].
  • 380 ± 180 ng μL−1 (24 h)−1 brussels sprout leaves infested by caterpillars [77].
  • 0, (2.5 ± 0.9) ng g DW−1 h−1 from intact and moth-damaged cabbage, resp. [81].
  • 0.6 ± 0.1 μg g−1 h−1 cut sugarcane in a chamber [48].
  • (55.5 ± 14.6 | 26.1 ± 20.1), (202 ± 68 | 62.7 ± 34.5), (8.8 ± 3.8 | -); (9.0 ± 15.2 | -) ng g DW−1 (June | August), Betula pendula, B. pubescens, Populus tremula, Sambucus nigra, resp., (dry leaves) [41].
  • (247.08 ± 172.86)/(27.18 ± 6.67/16.95 ±6.97)/(218.05 ± 110.23) ng gDW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed; resp.) [44].
  • (43.22 ± 14.97) / (190.16 ± 47.64) ng g DW−1 h−1 Betula pendula (control/aphid infested for 21 days) [44];
  • (14.42 ± 4.63) / (57.58 ± 31.53) ng g DW−1 h−1 Alnus glutinosa (control/aphid infested for 21 days) [44].
  • 0.1–13.9 μgC g DW−1 h−1 various trees (cut branches) [90].
  • 2.29–7.7 ng g (foliage)−1 h−1 five tree species [40].
(Z)-3-hexenol
+ (E)-3-hexenol
+ (E)-2-hexenol
  • 0.1 kg C ha−1 yr−1 from switchgrass plantations (growth and harvest);
    6 ± 3 ng g DW−1 h−1 (1 ± 0.5 pmol m−2 s−1) before cutting;
    466 ± 144 ng g DW−1 h−1 (133 ± 35 pmol m−2 s−1) after cutting;
    149 ± 14 ng g DW−1 h−1 (42 ± 4 pmol m−2 s−1) mean [88].
Hexanal
  • 7.3, (25.3, 2.1), 10.4, 26.9, 19.2, 7.3, (6.4, 4.8), 6.3, 22.2, 13.6, (1.0, 3.1, 2.1), and 4.6 ng g DW−1 h−1 grape, (grape with fruits), rape, blossoming rape, rye, blossoming rye, (beech), hornbeam, birch, oak, (grassland), grassland after mowing, resp. (in a field enclosure) [86].
  • 2 and 5 nmol g−1 fresh weight of intact and homogenized cucumber leaves [38].
  • 4.7 ± 5.3, 1.3 ± 3.0 ng g−1 Betula pendula, B. pubescens (dry leaves), June [41].
  • 0.6, 0.1 μgC g DW−1 h−1 rabbit brush, Populus tremuloides (cut branches) [90].
(Z)-2-hexenal
  • 4.25 ± 0.64 pmol m−2 s−1 L. pulmonaria thalli (intact);
    7.19 ± 1.98, 6.61 ± 0.16, 10.0 ± 1.9 pmol m−2 s−1 after heat shock of 37, 46, and 51 °C, resp.;
    8.84 ± 1.31 pmol m−2 s−1 (wounded) [87] b.
  • 0.2, 0.3, 0.5 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16].
(E)-2-hexen-1-al
  • 8.6 nmol m−2 s−1 mountain grassland, cut [96].
  • <0.4 and 9.8 pmol m−2 s−1 Bel-W3 tobacco, before and after exposure to O3 in a lab reactor [89].
  • 8.7, 0.8, 8.9 ng g DW−1 h−1 oak, grassland, and grassland after mowing (in a field enclosure) [86].
  • 40, 5 nmol m−2 s−1 from Phragmite leaves exposed to 45 °C in normal and O2-deprived atmosphere;
    5, 6.8 nmol m−2 s−1 from Phragmite leaves and from Arabidopsis NPQ1 plants exposed to high light, resp. [14].
  • 8, 28 nmol g−1 fresh weight of intact and homogenized cucumber leaves [38].
  • (8.6 ± 11.1) / (0.7 ± 1.4) ng g DW−1 Betula pendula dry leaves, June/August [41].
(Z)-3-hexen-1-al
  • 0.5 mg C g−1 of drying aspen leaves [33].
  • 50 μgC g−1 h−1 cut and drying Festuca rubra (red fescue) [45].
  • 1.6 and 49.3 pmol m−2 s−1 Bel-W3 tobacco, before and after exposure to O3 in a lab reactor [89].
  • 356 ± 95, 4872 ± 633 ng/30 min from caterpillar-infested corn plant and its cut leaves, resp. (in a chamber) [34].
  • 2500 nmol g−1 fresh weight of homogenized Arabidopsis leaves [39].
  • 25, 41 nmol g−1 fresh weight of intact and homogenized cucumber leaves, resp. [38].
(Z)-3-hexenal
+ (E)-2-hexenal
  • 0.2 kg C ha−1 yr−1 from switchgrass plantations (growth and harvest);
    4 ± 6 ng g DW−1 h−1 (0.6 ± 2 pmol m−2 s−1) mean before cutting;
    2666 ± 4787 ng g DW−1 h−1 (731 ± 1315 pmol m−2 s−1) max after cutting;
    486 ± 903 ng g DW−1 h−1 (133 ± 247 pmol m−2 s−1) mean after cutting [88].
Hexyl acetate
  • (14.4 ± 4.0 | 2.3 ± 4.5), (47.4 ± 13.6 | 13.9 ± 6.7) ng g DM−1 (June | August), Betula pendula, B. pubescens, resp. (dry leaves) [41].
  • 0.43–3.16 ng g (foliage)−1 h−1 five tree species [40].
(Z)-3-hexenyl acetate
  • 8.66 ± 0.08 μg m−2 mowed lawn [92].
  • <0.3 and 39.6 pmol m−2 s−1 Bel-W3 tobacco, before and after exposure to O3 in a lab reactor [89].
  • 0.21 ± 0.06, 0.27 ± 0.02 pmol m−2 s−1 intact apple and grape foliage day mean, resp. [93].
  • 595 ± 569, 3720 ± 1052 ng/30 min from caterpillar-infested corn plant and its cut leaves, resp. (in a chamber) [34].
  • 0.1, 0.2, 2 nmol m−2 s−1 Couepia longipendula Pilg. (detached leaves over Amazon Forest canopy, February, peak values at 7:30, 12:40, 16:40, resp.) [16].
  • 10.9, (114.1, 9.4), 37.5, 117.3, 11.1, (3.4, 4.1), 50.9, 95.0, 5.4, (1.8, 13.8, 13.1), and 29.6 ng g DW−1 h−1 grape, (grape with fruits), rape, blossoming rape, blossoming rye, (beech), hornbeam, birch, oak, (grassland), grassland after mowing, resp. (in a field enclosure) [86].
  • 0.7, 1.9, 1.1, 0.4, ≤ 0.2, (0.8, 0.5), 0.1, 0.6, 0.2, 0.06, 3.4, 1.6, 0.4, 0.3, 0.2, and 3.3 μg g DW−1 h−1 alfalfa, almond, apricot, cherry, cotton (grape), lemon, nectarine, olive, orange, peach, plum, sorghum, walnut, valley oak, and whitehorn, resp. (in flow-through field enclosures) [95].
  • 3608 ± 2076, 490 ± 220, 340 ± 200 ng μL−1 (24 h)−1 brussel sprout leaves infested by caterpillars, non-infested artificially damaged leaves from the infested plant, and undamaged leaves, resp. [77].
  • 0, 8.4 ± 3.8 ng g DW−1 hr−1 from intact and moth-damaged cabbage, resp. [81].
  • 1.22 ± 0.18 μg g−1 hr−1 cut sugarcane in a chamber [48].
  • 7.35 ± 1.02 pmol m−2 s−1 L. pulmonaria thalli (intact);
    7.36 ± 1.41, 12.1 ± 1.1, 17.5 ± 2.5 pmol m−2 s−1 after heat shock: 37 °C, 46 °C, 51 °C;
    20.5 ± 1.4 pmol m−2 s−1 (wounded) [87].
  • (130 ± 42.8/44.4 ± 40.0), (501 ± 246 / 26.4 ± 16.3), (67.6 ± 18.6/-), (14.3 ± 6.0/-ng) g DW−1 Betula pendula, Populus tremula, B. pubescens, Sambucus nigra, resp. (dry leaves), June/August [41].
  • (41.18 ± 20.87)/(8.96 ± 1.61)/(6.90 ± 2.38)/(34.89 ± 13.66) ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed) [44].
  • (15.99 ± 5.08)/(62.38 ± 16.54) ng g DW−1 h−1 Betula pendula (control/aphid infested for 21 days) [44].
  • (21.44 ± 6.11)/(43.91 ± 20.52) ng g DW−1 h−1 Alnus glutinosa (control/aphid infested for 21 days) [44].
  • 0.3–19 μgC g DW−1 h−1 various trees (cut branches) [90].
  • 6.28–71.39 ng g (foliage)−1 h−1 five tree species [40].
  • (75.5 ± 5.7) − (178.1 ± 31.3) ng g DW−1 h−1 Pinus sylvestris (various seedlings) [91].
(E)-2-hexenyl acetate
  • (4.0 ± 2.1 / 0.4 ± 0.8); (38.7 ± 22.2/-) ng g−1 Betula pendula, B. pubescens, resp. (dry leaves), June/August [41].
  • 0.1, 0.2 μg C h−1 g DW−1 ironwood (Accacia), slippery elm (Ulmus), resp. [90].
  • 0.12 ± 0.07, 0.35 ± 0.17 ng g (foliage)−1 h−1 blue ash, white fingertree [40].
(Z)-3-Hexenyl acetate
+ (E)-2-hexenyl acetate
  • 3 ± 2 ng g DW−1 h−1 (0.5 ± 0.3 pmol m−2 s−1) switchgrass, before cutting;
    35 ± 50 ng g DW−1 h−1 (6 ± 8 pmol m−2 s−1) after cutting;
    13 ± 18 ng g DW−1 h−1 (2 ± 3 pmol m−2 s−1) mean [88].
(Z)-3-hexenyl
butanoate
(butyrate)
  • 0.0306 ± 0.0008 pmol m−2 s−1 intact apple foliage day mean [93].
  • (-| 20.8 ± 10.1), (0.7 ± 1.1 | -) ng g DW−1 (June | August), Betula pubescens, Populus tremula, resp. (dry leaves) [41].
  • (12.84 ± 8.86)/(0.48 ± 0.32)/(0.37 ± 0.37)/(4.45 ± 1.71) ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed) [44].
  • (0.19 ± 0.19)/(2.42 ± 1.04) ng g DW−1 h−1 Betula pendula
    (control/aphid infested for 21 days) [44].
  • 0.2–5.2 μg C g DW−1 h−1 various trees (cut branches) [90].
  • (0.9 ± 0.9) − (1.3 ± 0.9) ng g DW−1 h−1 Pinus sylvestris (various seedlings) [91].
(E)-3-hexenyl
butanoate
  • (7.3 ± 2.1 | -), (33.8 ± 11.3 | 7.3 ± 8.5), (4.8 ± 1.4 | -) ng g DW−1 (June | August), Betula pendula, B. pubescens, Populus tremula, resp. (dry leaves) [41].
(Z)-3-hexenyl 3-methylbutanoate
(isovalerate)
  • (0.46 ± 0.46)/(0.30 ± 0.30)/(0 ± 0)/(5.50 ± 1.70) ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed) [44].
  • 0.49 ± 0.49 ng g DW−1 h−1 Betula pendula [44].
  • 0.3, 0.8 μg C g DW−1 h−1 Artemisa tridentata, Betula alleghaniensis (cut branches) [90].
(Z)-3-hexenyl 2-
methyl-2-butenoate
(tiglate)
  • (1.21 ± 0.54)/0.13 ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days [44].
3-hexenyl hexanoate c
  • 2.9 ± 2.3 ng g DW−1 Populus tremula (leaves), June [41].
(Z)-3-hexenyl
isobutyrate
  • 0.03 ± 0.01, 0.40 ± 0.15, 0.03 ± 0.01 ng g−1 h−1 black ash, white fingertree, and olive tree, resp. (foliage) [40].
(Z)-3-hexenyl
butyrate
  • 0.15–0.66 ng g−1 h−1 five tree species (foliage) [40].
(Z)-3-hexenyl
isovalerate
  • 0.06–1.34 ng g−1 h−1 five tree species (foliage) [40].
(Z)-3-hexenyl isotiglate
  • 0.02–0.38 ng g−1 h−1 four tree species (foliage) [40].
hexenyl type
  • 140, 50, 0.06 μg (C) g DW−1 h−1 grass, clover (both cut and slashed), and over mowed grassland, resp [18].
1-octen-3-ol
  • 9.0 ng g DW−1 h−1 blossoming rye in a field enclosure [86].
  • (1.0 ± 0.6) − (44.6 ± 40.4) ng g DW−1 h−1 Pinus sylvestris (various seedlings) [91].
2-methyl-3-buten-2-ol
  • 3.93 ± 1.70, 2.69 ± 1.24 nmol m−2 s−1 Pinus ponderosa, drought-stressed and watered [97].
  • 6.933, 0.522 nmol m−2 s−1 mean, Pinus ponderosa and Pinus nigra, de novo synthesis [98].
  • 1.6, 0.77 nmol m−2 s−1 (max, daily mean, resp.) Pica abies flux over mountain forest [99].
  • 0.4–1.5 mg m−2 h−1 over mixed coniferous Roosevelt National Forest, Colorado (the lowest and peak values) [46].
  • 76, 530 μg C m−2 h−1 Pinus taeda (plantation canopy level), mid-spring to summer and fall to early spring, resp. [47].
  • 1.1 mg C m−2 h−1 Pinus ponderosa (plantation canopy level average), July [100].
  • 1.37/0.01 mg C m−2 h−1 Pinus ponderosa (plantation average canopy level), daytime/nighttime, July to September [101].
  • 1.0 mg C m−2 h−1 subalpine forest in the Rocky Mountains (canopy level average), July–August [102].
  • 0.3–1.4 mg C m−2 h−1 Pinus taeda (canopy level, plantation near Chappel Hill) July [103].
  • 2 mg C m−2 h−1 Pinus ponderosa (plantation canopy level average), July August;
    18 μg C g−1 h−1 (max. at leaf level), at 30 °C and 1000 mmol m−2 s−1 PAR [104].
  • 18 μg ± 7.4 C g−1 h−1 mean Pinus ponderosa at 30 °C and high irradiance [98].
  • 25 μg C g−1 h−1 average Pinus ponderosa 1-year-old needles, intact [105].
  • Over 20 mg C g−1 h−1 from leaves exposed to sunlight at 30 °C [45].
  • 0.2, 0.3 μg C g DW−1 h−1 Picea engelmanni, Pinus contorta (cut branches) [90].
Nonanal
  • (12.94 ± 0.95)/(12.57 ± 1.78)/(23.50 ± 3.09)/(25.12 ± 2.53) ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed) [44].
  • (6.51 ± 1.61)/(3.94 ± 2.76) ng g DW−1 h−1 Betula pendula (control/aphid infested for 21 days) [44].
  • (3.92 ± 0.87)/(0.79 ± 0.50) ng g DW−1 h−1 Alnus glutinosa (control/aphid infested for 21 days) [44].
Methyl salicylate
  • Up to 0.25 mg m−2 h−1 over walnut agroforest after cold night (7.5 °C) [106].
  • <0.2 and 123 pmol m−2 s−1 tobacco Bel-W3 before and after exposure to O3 in a lab reactor [89].
  • 40 ± 17 pmol m−2 s−1 from drought-stressed Alnus glutinosa in a chamber;
    63 ± 17 pmol m−2 s−1 with additional insect treatment;
    14 ± 5 pmol m−2 s−1 after re-watering [107].
  • 2.8 ng cm−2 h−1 from aphid-infested Pinus sylvestris saplings [75].
  • (26.3 ± 22.0 | 1.2 ± 1.7), (9.2 ± 11.8 | -), (14.9 ± 3.8 |-), (0.7 ± 1.7 | -) ng g DW−1, (June | August), Betula pendula, B. pubescens, Populus tremula, Sambucus nigra, resp. (dry leaves) [41].
  • (25.45 ± 10.08)/(370.82 ± 99.35)/(551.32± 99.35)/(387.82 ± 77.14) ng g DW−1 h−1 Betula pendula (control/aphid infested for 10 days/aphids removed/washed) [44].
  • (172.09 ± 120.115)/(721.63 ± 204.91) ng gDW−1 h−1 Betula pendula (control/aphid infested for 21 days) [44].
  • (1.39 ± 1.39)/(137.98 ± 45.16) ng g DW−1 h−1 Alnus glutinosa (control/aphid infested for 21 days) [44].
  • 0.5, 0.3 μg C g DW−1 h−1 Fagus grandfolia, Quercus velutina (cut branches) [90].
  • 0.27–2.76 ng g−1 h−1 five tree species (foliage) [40].
Hexenals
  • 0.07–0.417 μmol g DW−1 various cut grasses [108].
  • 4 × 10−7 g m−2 s−1 hay after cutting and drying [109].
  • 100–240 μg g DW−1 hay drying in a cuvette;
    40–50 mg m−2 h−1 hay cutting in a chamber;
    0.2–0.4 mg m−2 h−1 hay drying in a chamber;
    0.1–1.5 mg m−2 h−1 1st day of hay drying on a pasture in St. Johann, Tirol [110].
Pentenols and 2-
methyl butanal
  • 15 μg g DW−1 hay drying in a cuvette;
    0.1–0.9 mg m−2 h−1 hay cutting in the chamber;
    0.07–0.1 mg m−2 h−1 hay drying in the chamber;
    0.1–0.7 mg m−2 h−1 1st day of hay drying on a pasture in St. Johann, Tirol [110].
Pentenols and 3-
methyl butanal
  • 8 × 10−8 g m−2 s−1 hay cutting and drying [109].
Hexanal and (E)-2-hexenol
  • 8 × 10−8 g m−2 s−1 hay cutting and drying [109].
Hexenols and hexanal
  • 0.07–0.211 μmol g DW−1 various cut grasses [108].
  • 30–60 μg g DW−1 hay drying in a cuvette;
    1–3 mg m−2 h−1 hay cutting in the chamber;
    0.1–0.3 mg m−2 h−1 hay drying in the chamber;
    0.1–0.3 mg m−2 h−1 1st day of hay drying on a pasture in St. Johann, Tirol [110].
Hexenyl acetates
  • 0.10–0.844 μmol g DW−1 various cut grasses [108].
Hexyl acetates
  • 0.003–0.015 μmol g DW−1 various cut grasses [108].
  • 2.71 ± 0.05, 0.29 ± 0.07 pmol m−2 s−1 intact apple and grape foliage day mean, resp. [93].
a leaf (or plant material) dry weight (mass); b isoprene fragment was observed pre-cut and more 1-penten-3-ol fragment post-cut on MS; c unspecified isomer.
Table 3. Ambient and headspace concentrations of GLV.
Table 3. Ambient and headspace concentrations of GLV.
GLVConcentrations and References
1-penten-3-ol
  • 17–54 ng dm−3 over freshly harvested tall fescue of various types in a headspace chamber [28].
(E)-2-penten-1-ol
  • 236–1529 ng dm−3 over freshly harvested tall fescue of various types (in a headspace chamber);
    1–9 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
(E)-2-pentenal
  • 1–4 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
1-penten-3-one
  • 7–55 ng dm−3 over freshly harvested tall fescue of various types (in a headspace chamber);
    3–9 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
1-hexanol
  • 1 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
  • 0.04–0.12 ppbv at the canopy level (30 m) October-September 2010 in Amazon Rainforest [16].
(E)-2-hexen-1-ol
  • 11–42 ng dm−3 over freshly harvested tall fescue of various types (in a headspace chamber);
    0–1 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
(Z)-3-hexen-1-ol
  • 958–3176 ng dm−3 over freshly harvested tall fescue of various types (in a headspace chamber);
    2–5 ng dm−3 over sun-cured hay of tall fescue of various types (in a headspace chamber) [28].
  • Max ~1000 ppbv, during cutting and drying grass and clover;
    10–100 ppbv from red fescue grass;
    5–100 ppbv from white clover leaves and stems, when cut and drying [12].
  • 732 pptv mixing ratio over the canopy of tropical rainforest in Surinam [65].
(E)-3-hexenol
  • 3–31 ng dm−3 over freshly harvested tall fescue of various types (in a headspace chamber) [28].
Hexanal
  • 5–28 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber [28].
(E)-2-hexenal
  • 22–458 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber;
    4–12 ng dm−3 over sun-cured hay of tall fescue of various types, in a headspace chamber [28].
  • 0.50–0.78 ppbv in 24 L of air at charcoal production workplace [111].
(Z)-2-hexenal
  • 11–72 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber;
    0–1 ng dm−3 over sun-cured hay of tall fescue of various types, in a headspace chamber [28].
(E)-3-hexenal
  • 26–165 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber;
    1–2 ng dm−3 over sun-cured hay of tall fescue of various types, in a headspace chamber [28].
(Z)-3-hexenal
  • 18–698 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber;
    0–1 ng dm−3 over sun-cured hay of tall fescue of multiple types, in a headspace chamber [28].
  • <100 pptv, ~16–20 ppbv, before and after lawn mowing [112].
  • Max ~1000 ppbv, during cutting and drying grass and clover;
    10–1000 ppbv from red fescue grass;
    10–100 ppbv from white clover leaves and stems when cut and dried in a lab chamber [12].
  • Up to 650 ppbv from cut aspen leaves in a chamber [33].
  • Up to 160 ppbv over cut drying bluegrass leaves (25–30 °C) [29].
(E)-2-hexenyl acetate
  • 49–102 ng dm−3 over freshly harvested various types of tall fescue in a headspace chamber [28].
(Z)-3-hexenyl acetate
  • 10,195–24,816 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber;
    0–2 ng dm−3 over sun-cured hay of tall fescue of various types, in a headspace chamber [28].
  • 165.3–353.7 pptv on a spring day;
    213.3–790.2 pptv in spring night, over flowering orange tree [113].
  • 1–10 ppbv from red fescue grass, 0.5–9 ppbv from white clover leaves and stems, when cut and dried in a lab chamber [12].
(Z)-3-hexenyl propionate
  • 1–5 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber [28].
1-octen-3-ol
  • 6–18 ng dm−3 over freshly harvested tall fescue of various types, in a headspace chamber.
    1–2 ng dm−3 over sun-cured hay of tall fescue of various types, in a headspace chamber [28].
2-methyl-3-buten-2-ol
  • 19 ppbv max., 0.2–2 ppbv avg. Pinus ponderosa plantation canopy level [104].
  • 0–7 ppbv Pinus ponderosa plantation in the Sierra Nevada, canopy level, July to September [101].
  • 50–200 pptv day (mixing ratio) Niwot Ridge, Colorado, June 1991 [114].
Methyl salicylate
  • Mixing ratio up to 14 ppbv over walnut agroforest after cold night (7.5 °C) [106].
Table 4. Henry’s constant, 1-octanol/water partition coefficient (KOW), and 1-octanol/air partition coefficient (KOA) for GLV in water and aqueous solutions.
Table 4. Henry’s constant, 1-octanol/water partition coefficient (KOW), and 1-octanol/air partition coefficient (KOA) for GLV in water and aqueous solutions.
GLVHenry’s ConstantOctanol/Water Partition
Coefficient
Octanol/Air Partition
Coefficient
HTIRef.Log (KOW)Ref.Log (KOA)Ref.
M atm−1KM
pentan-1-ol 87.7 ± 8.92980[120] c1.5[122]4.7[123]
84.4 ± 7.0
46.1 ± 2.1
11.9 ± 0.3
3.8 ± 0.1
298
306
323
343
[124] a1.34[125]f 4.8[117] i
71.9 ± 3.6298[117] g1.3[117] h4.6[117] i
1-penten-3-ol 72.5 ± 40.6298 [117] g0.84[126]4.5[117] i
0.81[125]
1.1[117] h
(Z/E)-2-penten-1-ol 119.8 ± 342980[117] g1.1[117] h4.4[117] i
(E)-2-pentenal 17.3 ± 12.3298 [117] g1.1[117] h3.6[117] i
1-penten-3-one 28.5 ± 23.2298 [117] g0.9[117] h3.8[117] i
(Z)-2-pentenyl acetate 2.8 ± 0.8298 [117] g2.1[117] h3.8[117] i
n-hexan-1-ol 61.8 ± 16.32980[120] c1.84[125]5.2[123]
59.6 ± 2.8
32.7 ± 1.1
8.0 ± 0.3
2.53 ± 0.01
298
306
323
343
[124] a2.03[122]5.0[117] i
55.5 ± 3.5298[117] g1.8[117] h
n-hexan-1-al 3.2 ± 0.42950[127]1.78[122]4.4[123]
5 ± 1298 [128]1.8[117] hf 3.8, 3.9[117] i
5.2 ± 0.5
4.4 ± 0.4
3.4 ± 0.3
3.6 ± 0.2
3.2 ± 0.1
freshwater
25% seawater
50% seawater
75% seawater
100% seawater
[129]
4.5 ± 1.00[120] c
3.9 ± 1.0 [117] g
hexyl acetate 1.5 ± 0.22980[127]2.59[126]4.6[123]
1.4 ± 0.2 [117] g2.8[117] h4.5 f, 4.4[117] i
hexenol b252980[33]
(E/Z)-2-hexen-1-ol 94.4 ± 35.2 [117] g1.6[117] h4.8[117] i
(Z)-3-hexen-1-ol 113 ± 7.1
140 ± 18
132 ± 11
83.4 ± 8.3
62.7 ± 3.0
298


303
308
0
0.01
1
0
0
[130]1.6[117] h4.8[117] i
129.9 ± 65.6298 [117] g
(E)-2-hexenal 14.5 ± 1.72980[127]1.6[117] h4.3 f, 4.0[117] i
20[128,131]
13.6 ± 7.3 [117] g
(Z)-3-hexenal 6 [33]1.6[117] h3.7[117] i
4.1 ± 1.9 [117] g
hexenyl acetate b1 [33]
(E/Z)-2-hexenyl acetate 1.8 ± 0.7 [117] g2.6[117] h4.2[117] i
(Z)-3-hexenyl acetate 3.1 ± 0.42980[127]2.6[117] h4.2[117] i
3.62 ± 0.22
3.29 ± 1.1
2.32 ± 0.17
3.21 ± 0.17
2.56 ± 0.19
298


303
308
0
0.01
1
0
0
[130]
3.3 ± 2.4298 [117] g
(E)-3-hexenyl acetate 3.3 ± 0.42980[127]
3.3 ± 2.4 [117] g
(Z)-3-hexenyl-
propionate
2.4 ± 1.6 [117] g3.1[117] h4.6[117] i
(E/Z)-3-hexenyl butyrate 1.8 ± 1.1 [117] g3.6[117] h4.9[117] i
(Z)-3-hexenyl isobutyrate 1.3 ± 0.6 [117] g3.5[117] h4.9[117] i
(E)-2-hexenyl butanoate 1.2 ± 0.3 [117] g3.6[117] h4.9[117] i
(Z)-3-hexenyl 3-methylbutanoate 1.1 ± 0.6 [117] g4.0[117] h5.2[117] i
(Z)-3-hexenyl 2-methyl-2-butenoate 1.8 ± 0.6 [117] g3.9[117] h5.4[117] i
3-hexenyl hexanoate 1.0 ± 0.5 [117] g4.6[117] h5.7[117] i
2-methyl-3-buten-2-ol 73 ± 3
38 ± 7
48 ± 26
296 ± 20
40% wt H2SO4
55% wt H2SO4
[132]1.1[117] h4.2 f, 4.5[117] i
48 a
52.9 ± 5.1
2980
0
[119]
38.7 ± 2.2
21.8 ± 4.4
40.2 ± 5.4
31.7 ± 2.2
303
308
 
 
0.01
1
0
0
[130]
65 ± 3.53030[128,133]
62 ± 0.830322.7 mM KNO3 + 7.42 mM CaSO4[133]
59.8 ± 58.6 d [117] g
1-octen-3-ol 13.2>378 K0[120,134]2.6[117] h5.6[117] i
44.5 ± 1.7 [117] g
nonanal 1.3 ± 0.22980[120] c3.3[117] hf 4.8, 5.0[117] i
1.2 ± 0.1
0.51 ± 0.03
0.38 ± 0.03
298Freshwater
50% seawater
100% seawater
[129]
1.6 ± 0.4 [117] g
Jasmonic acid e 2.5[117] h9.7[117] i
methyl jasmonate e5081 ± 10032980[106]2.8[117] h7.5[117] i
8091 ± 1121
5454 ± 520
3869 ± 261
6716 ± 1272
4837 ± 272
298
298
298
303
308
0
0.01
1
0
 
[130]
methyl salicylate e33.5 ± 4.02980[106]2.6[117] h5.0 f, 6.3[117] i
37.9 ± 2.1
26.7 ± 3.4
20.1 ± 1.6
16.4 ± 0.9
10.0 ± 4.2
298
298
298
303
308
0.01
1
0
 
 
[130]
a recalculated from the original data; b unspecified structure; c mean of experimental values summarized by Sander 2015; d EPI suite could not estimate the value using group estimation method, so the difference between bond estimation value and VP/WSol resulted in large std. deviation; e the predicted KH values are unreasonable and therefore not included f Log KOA (octanol/air): estimated within EPI suite using one of the experimental values; g mean of bond est., group est., and VP/Wsol est. values by EPI suite (HENRYWIN v3.20) for which standard deviation is provided as error; h calculated with EPI suite HENRYWIN v3.20, Log KOW (version 1.68 estimate); i calculated with EPI suite HENRYWIN v3.20, Log Octanol-Air (KOAWIN v1.10).
Table 5. Solubility in water and vapor pressure of GLVs at 298 K, experimental and estimated with EPI suite (MPBPVP (v1.43), mean of Antoine and Grain methods, and WSKOW v1.42 method, resp.).
Table 5. Solubility in water and vapor pressure of GLVs at 298 K, experimental and estimated with EPI suite (MPBPVP (v1.43), mean of Antoine and Grain methods, and WSKOW v1.42 method, resp.).
GLVSolubilityVapor Pressure
mmol dm−3Ref.AtmRef.
1-pentanol 245.5[136]2.9 × 10−3[137]
307[138]3.7 × 10−3[138]
257[126]3.49 × 10−3[117] b
236.98[117] a
1-penten-3-ol 1047.1[126]1.20 × 10−2[117] b
1035.1[125]
525.48[117] a
(Z/E)-2-penten-1-ol 530.83[117] a3.46 × 10−3[117] b
(E)-2-pentenal 178.08[117] a2.43 × 10−2[117] b
1-penten-3-one 260.34[117] a5.03 × 10−2[117] b
(Z)-2-pentenyl acetate 11.23[117] a4.14 × 10−3[117] b
n-hexan-1-ol 57.5[136]1.22 × 10−3[139]
61.3[125]1.16 × 10−3[117] b
67.39[117] a
n-hexanal 56.3 c[117] b1.49 × 10−2[139]
49.9[140]1.39 × 10−2[140]
35.21[117] a1.26 × 10−2[117] b
hexyl acetate 3.5[141]1.74 × 10−3[142]
8.9[126]1.91 × 10−3[117] b
2.14[117] a
hexenol f159.74[117] a
(E/Z)-2-hexen-1-ol 159.74[117] a1.20 × 10−3[117] b
(Z)-3-hexen-1-ol 162 ± 6[130]1.23 × 10−3[117] b
159.74[117]
(E)-2-hexenal 53.61[117] a6.21 × 10−3[117] b
(Z)-3-hexenal 53.61[117] a1.97 × 10−2[117] b
(E/Z)-2-hexenyl acetate 3.38[117] a2.61 × 10−3[117] b
(Z)-3-hexenyl acetate 3.12 ± 0.17[130]1.50 × 10−3[117] b
3.38[117] a
(E)-3-hexenyl acetate 3.38[117] a1.50 × 10−3[117] b
(Z)-3-hexenyl-propionate 1.02[117] a5.50 × 10−4[117] b
(E/Z)-3-hexenyl butyrate 0.31[117] a2.05 × 10−4[117] b
(Z)-3-hexenyl isobutyrate 0.35[117] a3.71 × 10−4[117] b
(E)-2-hexenyl butanoate 0.31[117] a2.05 × 10−4[117] b
(Z)-3-hexenyl 3-methylbutanoate 0.11[117] a1.39 × 10−4[117] b
(Z)-3-hexenyl 2-methyl-2-
butenoate
0.13[117] a7.53 × 10−5[117] b
3-hexenyl hexanoate 0.03[117] a3.04 × 10−5[117] b
2-methyl-3-buten-2-ol 1959 ± 36[130]3.08 × 10−2[117] b
565.89[117] a
1-octen-3-ol 14.32[117] a3.13 × 10−4[117] b
nonanal 0.7[140]4.87 × 10−4[139]
0.93[117] a5.13 × 10−4[140]
7.42 × 10−4[117] b
Jasmonic acid 3.53[117] a2.41 × 10−8 d
1.79 × 10−7 e
[117] b
methyl jasmonate 4.52 ± 0.09[130]4.43 × 10−7 d
1.24 × 10−6 e
[117] b
0.64[117] a
methyl salicylate 5.11 ± 0.06[130]7.03 × 10−5[117] b
4.6 c[141]
12.32[117] a
a estimated with EPI suite (WSKOW v1.42); b estimated with EPI suite (MPBPVP (v1.43), mean of Antoine and Grain methods); c 308 K; d selected VP with modified grain method; e subcooled liquid VP with modified grain method; f unspecified structure.
Table 6. Theoretical rate constants (cm3 molecule−1 s−1) for reactions of hexenols with OH radicals through the OH addition and hydrogen abstraction channels [171].
Table 6. Theoretical rate constants (cm3 molecule−1 s−1) for reactions of hexenols with OH radicals through the OH addition and hydrogen abstraction channels [171].
ReactantkOH additionkH abstractionktotalH-Abstraction
Branching Ratio
ktotal Deviation from
Experimental Values %
(E)-2-hexen-1-ol4.97 × 10−119.76 × 10−125.95 × 10−110.1626.4
(Z)-2-hexen-1-ol5.34 × 10−112.19 × 10−117.53 × 10−110.2911.7
(E)-3-hexen-1-ol2.45 × 10−113.50 × 10−115.95 × 10−110.5934.6
(Z)-3-hexen-1-ol1.17 × 10−103.07 × 10−111.48 × 10−100.21−46.5
(E)-4-hexen-1-ol4.74 × 10−111.30 × 10−116.04 × 10−110.2215.4
(Z)-4-hexen-1-ol1.10 × 10−102.73 × 10−111.37 × 10−100.2082.6
Table 9. Products observed in reactions of GLV with NO3.
Table 9. Products observed in reactions of GLV with NO3.
GLVProductYield, %
Molar
Expt
TabLe 10
Ref.
1-penten-3-ol,
(Z)-2-penten-1-ol
2-pentenyl nitrateobservedI a[192]
Ethyl vinyl nitrateobserved
2-penten-1-olobservedI b
(E)-2-hexenalPAN analogue100 ± 2.5II[193]
CO6.4 ± 4.2
Formic acidobserved
MBOAcetone63III[191]
68.7 ± 7.1IV[182]
1-nitroxyacetaldehyde67III[191]
Formaldehyde4III[191]
Organic nitrates12.5–13.5IV[182]
Hydroxy-aldehyde nitrates,
peroxynitrates, carbonyl nitrates, nitroxyalcohols, dinitrates
observed
Table 10. Experiments used to determine the rate constants listed in Table 9.
Table 10. Experiments used to determine the rate constants listed in Table 9.
ExptPhotoreactorDetectionTpNO3 SourceRef.
TypeVol. dm3Material Katm
I aChamber80TeflonGC-MS2981N2O5[192]
I bNO2
IIReactor977PyrexLP FTIR, UV2941N2O5[193]
IIIReactor153GlassLP FTIR2971N2O5[191]
IVChamber480TeflonLP FTIR2950.97N2O5[182]
Table 11. Products observed in reactions of GLV with ozone.
Table 11. Products observed in reactions of GLV with ozone.
GLVProductYield, MolarExpt
Table 12
Ref.
1-penten-3-olFormaldehyde0.49 ± 0.02I[201]
0.34 ± 0.04IX[199,200]
2-hydroxybutanal0.46 ± 0.03I[201]
0.30 ± 0.05 aIX[199,200]
Propanal0.15 ± 0.02I[201]
0.12 ± 0.01IX[199,200]
(Z)-2-penten-1-olPropanal0.39 ± 0.03I[201]
0.493 ± 0.075V[183]
0.51 ± 0.02IX[199,200]
3-hydroxypropanal0.33 + 0.33/−0.16V[183]
Glycolaldehyde0.43 ± 0.04I[201]
2-hydroxyacetaldehyde b0.53 ± 0.04IX[199,200]
Formaldehyde0.07 ± 0.02IX[199,200]
Acetaldehyde0.08 ± 0.01IX[199,200]
methylglyoxal0.04 ± 0.01IX[199,200]
1-penten-3-oneFormaldehyde0.37 ± 0.02I[201]
2-oxobutanal0.49 ± 0.03I[201]
SOA0.13–0.17I[201]
(Z)-3-hexen-1-olPropanalObservedII[92]
0.43 ± 0.02VIII[202]
0.59IX[199,200]
Propanoic acidObservedII[92]
2-propenalObservedII
AcetaldehydeObservedII
Ethane0.069 ± 0.005VIII[202]
OH0.26VI[174]
0.28 ± 0.06VIII[202]
(E)-2-hexenalOH0.62VI[174]
Butanal0.53IX[199,200]
0.527 ± 0.055IX[203]
Glyoxal0.56 [199,200]
0.559 ± 0.037IX[203]
2-oxobutanal c0.074 ± 0.006
acetaldehyde0.109 ± 0.020
Propanal0.067 ± 0.008
cyclohexanone0.032 ± 0.003
(Z)-3-hexenalPropanal0.35 ± 0.01VIII[202]
Ethane0.079 ± 0.004
OH0.32 ± 0.07
(Z)-3-hexenyl acetatePropanalObservedII[92]
0.76 ± 0.04
Propanoic acidObservedII[92]
2-propenalObservedII
Acetic acidObservedII
OH0.16VI[174]
acetaldehyde0.05 ± 0.01IX[199,200]
methylglyoxal0.05 ± 0.01
Acetone0.04 ± 0.01
glyoxal0.003 ± 0.001
(E)-2-hexenyl acetatebutanal0.47IX[199,200]
0.473 ± 0.023IX[203]
1-oxyethyl acetate0.58IX[199,200]
0.583 ± 0.141 cIX[203]
Glyoxal0.209 ± 0.004
Propanal0.102 ± 0.002
2-oxobutanal c0.091 ± 0.007
Acetaldehyde0.039 ± 0.003
Cyclohexanone0.090 ± 0.004
3-metyl-2-buten-3-olAcetone0.083 ± 0.050III[182]
ObservedIV[178]
0.67 ± 0.05VII a[177]
0.23 ± 0.06IX[199,200]
Formaldehyde0.467 ± 0.055III[182]
0.29 ± 0.03IV[178]
0.44 ± 0.05VII a[177]
0.55 ± 0.03VII b[177]
0.36 ± 0.09IX[199,200]
Formic acidObservedIII[182]
0.14 ± 0.04VII a[177]
Formic anhydride0.16 ± 0.07
2-hydroxy-2-methylpropanal0.30 ± 0.060.47 (FTIR)IV[178]
0.43 ± 0.12VII a[177]
0.84 ± 0.08VII b[177]
0.30 ± 0.02IX[199,200]
Acetaldehyde0.01 ± 0.01
CO0.343 ± 0.024III[182]
CO20.232 ± 0.015III
OH0.19IV[178]
a measured as 2-oxobutanal; b measured as glyoxal; c tentative.
Table 12. Experiments used to determine the rate constants listed in Table 11.
Table 12. Experiments used to determine the rate constants listed in Table 11.
ExptPhotoreactorDetectionTemp.pAdditionsRef.
TypeVol. dm3Material Katm
ICRAC
LISA
3910
977
FEP
Pyrex
LP FTIR293 b
295 c
1CO a[201]
IIChamber775TeflonGC-MS2961 [92]
IIIChamber480TeflonLP FTIR2950.97 [182]
IVChamber7000–8000, 5800TeflonAPI-MS/MS,
GC-FID,
GC-MS,
GC-FTIR,
LP_FTIR
2980.97 [178]
VChamber6500–7900TeflonGC-FID, GC-MS298 Cyclohexane a[183]
VIChamber6700,
7900
TeflonGC-FID2960.97Cyclohexane a
(E)-2-butene d
[174]
VII aLISA
CRAC
980
3910
Pyrex
FEP
LP-FTIR
GC-MS
298 e1CO a[177]
VII bCRAC3910FEPLP-FTIR
GC-MS
298 f1CO a[177]
VIIIChamber6000 LP FTIR29811,3,5-trimethyl
benzene a
[202]
IXChamber3700–3900TeflonLC-UV282–2961Cyclohexane a[199,200]
a OH scavenger; b RH < 0.4%; c RH < 1%; d reference reactant; e RH <0.2%; f RH ≈30%.
Table 13. Products from gas-phase reactions of GLV with Cl radicals.
Table 13. Products from gas-phase reactions of GLV with Cl radicals.
GLVProductYield %Expt
Table 14
Ref.
1-penten-3-olChloroacetaldehyde
Propanal
Acetaldehyde
1-penten-3-one
33 ± 1
39 ± 1
8 ± 3
<2
I[205]
(Z)-2-penten-1-ol2-chlorobutanal
Propanal
Acetaldehyde
(Z)-2-pentenal
19 ± 1
27 ± 1
18 ± 2
36 ± 1
I[205]
(E)-2-hexenyl acetatebutanalnot quantifiedV[186]
formaldehyde
propanal
2-methyl-3-buten-2-olAcetone47 ± 5II[179]
48 ± 4III[207]
38.5 ± 20.6IV[204]
chloroacetaldehyde53 ± 5II[179]
47 ± 5III[207]
formyl chloride<11II[179]
6.8 ± 5.9IV[204]
formaldehyde6 ± 2II[179]
7.2 ± 0.6III[207]
glycolaldehyde2.1 ± 0.2III[207]
formic acidBelow the quantification limitIII[207]
1.8 ± 0.1IV[204]
HCl15.7 ± 13.5IV[204]
CO20 ± 9.3
CO2<15
Table 14. Experiments used to identify the products listed in Table 13.
Table 14. Experiments used to identify the products listed in Table 13.
ExptPhotoreactorDetectionTemp.pCl SourceRef.
TypeVol. dm3Material KAtm
IBag400TeflonGC-FID, GC-MS2981Trichloroacetyl chloride[205]
IIChamber47SteelLP FTIR295 Cl2[179]
IIIChamber140PyrexFTIR296 ± 20.92Cl2[207]
IVChamber480TeflonFTIR298 ± 20.97Cl2[204]
VBag80TeflonGC-FID,
GC-MS
(SPME)
2981oxalyl chloride[186]
Table 15. Relative and absolute (a) rate constants for gas-phase reactions of GLV with OH at single temperatures.
Table 15. Relative and absolute (a) rate constants for gas-phase reactions of GLV with OH at single temperatures.
GLVk
cm3 molecule−1 s−1
T
K
P
Atm
Expt
Table 16
Ref.
Pentan-1-ol(1.0 ± 0.1) × 10−112981I[210]
(1.08 ± 0.11) × 10−11 a2960.03–0.07II[211]
(1.20 ± 0.16) × 10−11 a298 ± 21III[212]
(1.05 ± 0.13) × 10−11ibidIV
(1.11 ± 0.11) × 10−11298 ± 2740 mmHgV[184]
(1.23 ± 0.10) × 10−11295 ± 21VI[213]
(1.26 ± 0.07) × 10−11
(1.20 ± 0.06) × 10−11
298 ± 21VII[214]
1-penten-3-ol(6.7 ± 0.9) × 10−112981VIII[185]
(Z)-2-penten-1-ol(1.06 ± 0.15) × 10−10
(E)-2-pentenal(2.35 ± 0.21) × 10−11 a2980.132–0.526IX[85]
1-penten-3-one3.6 × 10−11 b2981X[215]
Hexan-1-ol(1.58 ± 0.35) × 10−11296 ± 21XI[216]
(Z)-2-hexen-1-ol(1.1 ± 0.4) × 10−10296 ± 2 XII[217]
0.95 × 10−10 b298
(8.53 ± 1.36) × 10−11 e2981XIII[171]
(E)-2-hexen-1-ol(1.0 ± 0.3) × 10−102981XIV[176]
(8.08 ± 1.33) × 10−11 e2981XIII[171]
(Z)-3-hexen-1-ol(1.2 ± 0.2) × 10−102981XIV[176]
(9.57 ± 2.42) × 10−11 a298 XV[57]
(1.08 ± 2.2) × 10−10296 ± 20.974XVI[174]
(1.01 ± 0.16) × 10−10 e2981XIII[171]
(E)-3-hexen-1-ol(1.4 ± 0.3) × 10−102981XIV[176]
(1.14 ± 0.14) × 10−10298 ± 21XVII[175]
(0.8 ± 0.1) × 10−10296 20.98 ± 0.01XII[217]
1.28 × 10−10298
(9.10 ± 1.50) × 10−11 e2981XIII[171]
(E)-4-hexen-1-ol(7.14 ± 1.20) × 10−11 e2981XIII[171]
(Z)-4-hexen-1-ol(7.86 ± 1.30) × 10−11 e
n-hexanal(2.60 ± 0.21) × 10−11 a2980.132–0.526IX[85]
(2.86 ± 0.13) × 10−11298 ± 21.00 ± 0.01XVIII[218]
(3.17 ± 0.15) × 10−11296 ± 21XIX[219]
(2.71 ± 0.20) × 10−11298 ± 21XX[220]
(E)-2-hexenal(0.681 ± 0.049) × 10−11296 ± 20.974XVI[174]
(2.95 ± 0.45) × 10−11 a2980.132–0.526IX[85]
(3.95 ± 0.17) × 10−11298 ± 21XXI[221]
(Z)−3-hexenal(6.9 ± 0.9) × 10−112981XXII[160]
(Z)- 3-hexenyl formate(4.61 ± 0.07) × 10−112981XXIII[222]
(4.24 ± 0.07) × 10−11 b
(E)-2-hexenylacetate(6.88 ± 1.41) × 10−11298 ± 1750 mmHgXXIV[186]
(Z)- 3-hexenyl acetate(1.21 ± 0.71) × 10−11296 ± 20.974XVI[174]
(Z)-3-hepten-1-ol(1.28 ± 0.23) × 10−10298 ± 21XVII[175]
2-methyl-3-buten-2-ol(6.9 ± 1.0) × 10−11295 ± 10.921XXV[179]
(5.67 ± 0.13) × 10−11296 ± 20.974XIX[223]
(3.9 ± 1.2) × 10−11298 ± 20.974 ± 0.007XXVI[224]
(6.6 ± 0.5) × 10−11298 ± 21XXVII[225]
(5.6 ± 0.6) × 10−11298 ± 21XXVIII[177]
(5.49 ± 0.44) × 10−11 a3005XXIX[226]
(6.32 ± 0.27) × 10−11 a,d
(6.61 ± 0.66) × 10−11 a,c
(Z)-3-octen-1-ol(1.49 ± 0.35) × 10−10298 ± 21XVII[175]
Nonanal(3.6 ± 0.7) × 10−102981XXX[188]
(2.88 ± 0.20) × 10−11298 ± 21XX[220]
Methyl salicylate(3.20 ± 0.46) × 10−122981XXXI[227]
a absolute rate constant; b from SAR or LFER; c reaction with OD; d with 15% of O2 added; e theoretical rate constants for the addition and hydrogen abstraction channels are given in Section 4.1.1 (Table 6).
Table 16. Experiments used to determine the rate constants listed in Table 15.
Table 16. Experiments used to determine the rate constants listed in Table 15.
ExptPhotoreactorDetectionOH SourceAdd.Reference
Reactants
Ref.
TypeVol. dm3Material
IBag140TeflonGC-FIDH2O2 Pentane
1,3-dioxolane
Cyclohexane
[210]
IIBulbs5PyrexRF H2O no[211]
IIICylinder50TeflonGC-FIDH2O2 or CH3ONO Cycloxexane[212]
IVCell1SteelLP UV no
VCylinder480DuranLP FTIR;
GC-PI
CH3ONONO aCycloxexane[184]
VIBag100TeflonGC-FIDH2O2 Propane;
hexane
[213]
VIIBag100TedlarGC-MSCH3ONO Cyclohexane;
p-xylene
[214]
VIIIChamber47SteelLP FTIRCH3CH2ONO propene[185]
IXCell PLP0.2PyrexLIFH2O2 no[85]
XSSR LFER [215]
XIChamber7600TeflonGC-FIDCH3ONONO aCyclohexane[216]
XIIBag200TeflonGC-FIDH2O2 Allyl ether
Cycloxexane
1-methyl-cyclohexane
[217]
XIIIBag150TeflonGC-FIDH2O2 Cyclohexane
2-methyl-3-buten-1-ol
[171]
XIVBag80TeflonGC-FID, SPMEH2O2 Methyl methacrylate
(E)-2-buten-1-ol
[176]
XVCell PLP0.2PyrexUV, LIFH2O2 no[57]
XVIChamber6700
7900
TeflonGC-FIDCH3ONONO a(E)-2-butene[174]
XVIIChamber1080QuartzLP FTIRH2O2 (E)-2-butene
isobutene
[175]
XVIIIReactor250SteelLP FTIRCH3ONO Other nitrites Propene
1-butene
[218]
XIXChamber7000TeflonGC-FIDCH3ONONO a1,3,5-trimethylbenzene Methyl vinyl ketone[219,223]
XXCylinder108OPyrexLP FTIRCH3ONO (E)-2-butene;
Methylvinyl ketone
[220]
XXIBag100TedlarGC-MSCH3ONO 2-methyl-2-butene, β-pinene[221]
XXIIChamber6000PFALP FTIRCH3ONO 1,3,5-trimethylbenzene[160]
XXIIIBag400TeflonGC-FIDH2O2 Cyclohexane,
n-octane, 1-butene
[222]
XXIVBag80TeflonGC-FIDH2O2 (E)-3-hexen-1-ol
2-buten-1-ol
[186]
XXVChamber47SteelLP FTIRCH3ONO Ethylene,
propylene
[179]
XXVIChamber480TeflonLP FTIRH2O2 Isoprene, propene[224]
XXVIIChamber6000TeflonLP FTIRCH3ONONO adi-n-butyl ether, propene[225]
XXVIIILISA
CRAC
EUPORE
980,
3910
2 × 105
Pyrex
FEP
Teflon
LP FTIRHONO
H2O2
NOxIsoprene[177]
XXIXDischarge-flow reactor0.02PyrexLIFF + H2O,
H + NO2
no[226]
XXXChamber5000TeflonGC-FIDIsopropyl
nitrite
NOxOctane[188]
XXXIICARE-CHRS chamber7300TeflonHR PTR-TOF-MSCH3ONO Methyl ethyl
ketone;
Toluene;
Di-n-butyl ether
[227]
a suppresses the O3 formation.
Table 17. Absolute (a) rate constants for gas-phase reactions of GLV with OH radicals—Arrhenius parameters and values at selected temperature T.
Table 17. Absolute (a) rate constants for gas-phase reactions of GLV with OH radicals—Arrhenius parameters and values at selected temperature T.
GLVAEA/RkTTT RangePExptRef.
cm3 Molecule−1 s−1Kcm3 Molecule−1 s−1KKAtmTable 18
Pentan-1-ol(6.7 ± 3.8) × 10−12 a−(132 ± 176)(1.04 ± 0.59) × 10−11 a298273–373100 mmHgI[210]
1-penten-3-ol(6.8 ± 0.7) × 10−12 a−(690 ± 20)(7.12 ± 0.73) × 10−11 a297243–40420–100 mmHgII[228]
(7.7 ± 1.6) × 10−12 a−(606 ± 60) a(5.65 ± 0.76) × 10−11 a298263–353 III[57]
(E)-2-penten-1-ol(6.8 ± 0.8) × 10−12 a−(680 ± 20)(6.76 ± 0.70) × 10−11 a
1-penten-3-one(4.4 ± 2.8) × 10−12 a−(507 ± 180) a(2.36 ± 0.47) × 10−11 a
(E)-2-pentenal(7.9 ± 1.2) × 10−12 a−(510 ± 20)(4.3 ± 0.6) × 10−11 a297244–3740.03–0.197IV[173]
(E)-2-hexen-1-ol(5.4 ± 0.6) × 10−12 a−(690 ± 20)(6.15 ± 0.75) × 10−11 a298263–353 III[57]
(Z)-3-hexen-1-ol(1.3 ± 0.1) × 10−11 a−(580 ± 10)(1.06 ±0.12) × 10−10 a297243–40420–100 mmHgII[228]
hexanal(4.2 ± 0.8) × 10−12 a−(565 ± 65)(2.78 ± 0.50) × 10−11 a298263–3530.066IV[159]
(E)-2-hexenal(7.5 ± 1.1) × 10−12 a−(520 ± 30)(4.4 ± 0.5) × 10−11 a297244–3740.03–0.197IV[173]
(9.8 ± 2.4) × 10−12 a−(455 ± 80)(4.68 ± 0.50) × 10−11 a298263–3530.066IV[159]
2-methyl-3-
buten-2-ol
(8.2 ± 1.2) × 10−12 a
8.43 × 10−12 b
−(610 ± 50)
−619.4 b
(5.4 ± 0.4) × 10−11 a
(6.4 ± 0.6) × 10−11 b
299 298254–410
254–360
1.3IV[165]
Methyl salicylate5.0962 × 10−11 a,c,d9171.9595 × 10−12 a,c,d298278–2981V[190]
1.54 × 10−12 a,d,e5053.96.64 × 10−20 a,d,e298278–3501VI[189]
a absolute rate constants; b values for reactions with OD radicals; c converted from the original units; d calculated theoretically; e initial OH addition, the paper contains Arrhenius equations for other steps of the reaction mechanism.
Table 18. Experiments used to determine the rate constants listed in Table 17.
Table 18. Experiments used to determine the rate constants listed in Table 17.
ExptPhotoreactorDetectionOH SourceRef.
TypeVol. dm3Material
ICell, LFP PyrexLIFH2O2[210]
IICell, LFP0.15PyrexLIFH2O2[173,228]
IIICell PLP0.2PyrexUV, LIFH2O2[57,159]
IVCell, PLP0.15PyrexLIFHONO,
DONO
[165]
VRKKM theory at MN15-L/aug-cc-pVTZ level with Eckart
tunneling correction
[190]
VIDFT theory using B3LYP, M06-2X, and MPW1K functionals with 6-311++G(d,p) basis set.[189]
Table 19. Relative and absolute (a) rate constants for gas-phase reactions of GLV with NO3 at single temperatures.
Table 19. Relative and absolute (a) rate constants for gas-phase reactions of GLV with NO3 at single temperatures.
GLVk
cm3 molecule−1 s−1
T
K
P
Atm
Expt.
Table 20
Ref.
1-penten-3-ol(1.39 ± 0.19) × 10−14 a298 ± 31.00 ± 0.03I a[192]
(Z)-2-penten-1-ol(1.53 ± 0.23) × 10−13 a
(3.5 ± 1.9) × 10−13I b
(3.11 ± 0.11) × 10−13295 ± 21II[230]
(E)-2-pentenal(1.93 ± 0.40) × 10−14
1-penten-3-one3.39 × 10−14 f2981III[215]
(Z)-4-hexen-1-ol(2.93 ± 0.58) × 10−13 a298 ± 3 I a[231]
(Z)-3-hexen-1-ol(2.67 ± 0.42) × 10−13 a
(2.72 ± 0.83) × 10−13296 ± 20.974IV[174]
(E)-3-hexen-1-ol(4.43 ± 0.91) × 10−13 a298 ± 3 I a[231]
(5.2 ± 1.8) × 10−13 gI b
(Z)-2-hexen-1-ol(1.56 ± 0.24) × 10−13 aI a
(4.05 ± 0.45) × 10−13 d
(3.57 ± 0.62) × 10−13 e
295 ± 21II[230]
(E)-2-hexen-1-ol(1.30 ± 0.24) × 10−13 a298 ± 3 I a[231]
hexanal(1.14 ± 0.14) × 10−14298 ± 21.00 ± 0.01V[218]
(1.73 ± 0.18) × 10−14298 ± 21V[232]
(4.75 ± 0.62) × 10−14 b
(1.07 ± 0.12) × 10−14 c
296 ± 21VI[219]
(1.7 ± 0.1) × 10−14297 ± 21VII a[233]
(Z)-2-hexenal(1.36 ± 0.29) × 10−14295 ± 21II[230]
(E)-2-hexenal(1.21 ± 0.10) × 10−14296 ± 20.974IV[174]
(4.7 ± 1.5) × 10−15 a294 ± 31VIII[193]
(Z)-3-hexenylacetate(2.46 ± 0.75) × 10−13296 ± 20.974IV[174]
2-methyl-3-buten-2-ol(2.1 ± 0.3) × 10−14 a2941IXa[234]
(1.6 ± 0.6) × 10−14IX b
(0.86 ± 0.29) × 10−14298 ± 20.974 ± 0.007X[224]
Nonanal(1.8 ± 0. 2) × 10−14
(2.2 ± 0. 3) × 10−14
297 ± 21VII b
VII c
[233]
Methyl salicylate(4.19 ± 0.92) × 10−152981XI[227]
a absolute rate constant; b methacrolein reference; c 1-butene reference, d (E)-2-butene reference, e cyclopentene reference; f estimated from SAR or LFER; g overestimated (see text).
Table 20. Experiments used to determine the rate constants listed in Table 19.
Table 20. Experiments used to determine the rate constants listed in Table 19.
ExptPhotoreactorDetectionNO3 SourceReference ReactantsRef.
TypeVol. dm3Material
I aDischarge flow tube Off-axis
continuous-wave CEAS
F + HONOno[192]
I bBag56TeflonGC-FIDN2O5(E)-2-butene,
1-butene
IIFlow-tube0.005GlassCI MSN2O5(E)-2-butene,
Cyclopentane,
1,3-butadiene
[230]
IIISRR, LFER [215]
IVChamber6700
7900
TeflonGC-FIDN2O5(E)-2-butene,
propene
[174]
VReactor250SteelLP FTIRN2O5Propene
1-butene
[218,232]
VIChanber6500TeflonGC-FIDN2O5Methacrolein,
1-butene
[219]
VII aBag100PVF(a) FTIRN2O5(a) propene[233]
VII b(b,c) GC (SPME)(b) butanal
VII c(c) 1-butene
VIIIReactor977PyrexFTIR, UVN2O5no[193]
IX aFlow tube FFD LP FTIRF + HNO3no[234]
IX bReactor153 N2O5propene
XChamber480TeflonLP FTIRN2O5propene[224]
XIICARE-CHRS chamber7300TeflonHR PTR-TOF-MSNO2 + O3acetaldehyde[227]
Table 21. Absolute rate constants for gas-phase reactions of GLV with NO3 radicals—Arrhenius parameters and values at temperature T.
Table 21. Absolute rate constants for gas-phase reactions of GLV with NO3 radicals—Arrhenius parameters and values at temperature T.
GLVAEA/RkTTT RangePExptRef.
Molecules cm−1 s−1KMolecules cm−1 s−1KKmmHgTable 22
(E) -2-pentenal(5.40 ± 0.30) × 10−121540 ± 200(2.88 ± 0.29) × 10−14 298298–4331I[235]
(E) -2-hexenal(1.2 ± 0.3) × 10−12926 ± 85(5.49 ± 0.95) × 10−14 298298–4331 [235]
2-methyl-3-butene-2-ol4.6 × 10−14400(1.2 ± 0.09) × 10−14298267–4001.2–8.8II[236]
Table 22. Experiments used to determine the rate constants listed in Table 21.
Table 22. Experiments used to determine the rate constants listed in Table 21.
ExpPhotoreactorDetectionOH SourceAdd.Reference
Reactants
Ref.
TypeVol. dm3Material
IFlow tube FFD LIFF + HNO3 no[235]
IIFlow tube LP VISN2O5 no[236]
III
IV
Table 23. Relative and absolute (a) rate constants for gas-phase reactions of GLV with O3 at single temperatures.
Table 23. Relative and absolute (a) rate constants for gas-phase reactions of GLV with O3 at single temperatures.
GLVk
cm3 molecule−1 s−1
T
K
P
atm
Expt
Table 24
References
pentan-1-ol(2.9 ± 0.2) × 10−172981 [210]
1-penten-3-ol(1.64 ± 0.15) × 10−17 a2981I[201]
(1.79 ± 0.18) × 10−17 a289 ± 11II[237]
(Z)-2-penten-1-ol(11.5 ± 0.66) × 10−17 a2981I[201]
(1.69 ± 0.25) × 10−16 a288 ± 11II[237]
1-penten-3-one(1.17 ± 0.15) × 10−17 a2981I[201]
(Z)-2-hexen-1-ol(7.44 ± 1.03) × 10−17 a2981III a[238]
10.4 × 10−17298 III b
(E)-2-hexen-1-ol(5.98 ± 0.73) × 10−172981IV [239]
(1.66 ± 0.22) × 10−16 a2981III a[238]
2.86 × 10−16298 III b
(Z)-3-hexen-1-ol6.7 × 10−172961V[92]
6.4 × 10−17
(1.05 ± 0.95) × 10−16 a258 II[237,240]
(6.4 ± 1.7) × 10−17296 ± 20.974VI[174]
(6.04 ± 0.07) × 10−172981IV [239]
(5.47 ± 0.71) × 10−17 a2981III a[238]
4.40 × 10−17298 III b
(E)-3-hexen-1-ol(5.83 ± 0.86) × 10−172981IV [239]
(6.19 ± 0.72) × 10−17 a2981III a[238]
5.8 × 10−17III b
(Z)-4-hexen-1-ol(7.09 ± 0.91) × 10−17 a2981III a[238]
8.38 × 10−17III b
(E)-4-hexen-1-ol(1.05 ± 0.14) × 10−16 aIII a
0.88 × 10−16III b
(E)-2-hexenal(1.28 ± 0.28) × 10−18287.3 ± 1.41II[203]
(3.0 ± 2.1) × 10−18296 ± 20.974VI[174]
(Z)-3-hexenal(3.45 ± 0.30) × 10−172981VII[160]
(Z)-3-hexenyl formate(4.06 ± 0.66) × 10−17 a2981VIII a[241]
6.60 × 10−17 a,bVIII b
(E)-2-hexenyl acetate(2.18 ± 0.28) × 10−17288.4 ± 0.41II[203]
(Z)-3-hexenyl acetate3.6 × 10−17; 5.4 × 10−172961V[92]
(5.4 ± 1.4) × 10−17296 ± 20.974VI[174]
(5.77 ± 0.70) × 10−17 a2981VIII a[241]
6.84 × 10−17 a,bVIII b
(Z)-3-hexenyl propionate(7.62 ± 0.88) × 10−17 a2981VIII a[241]
1.084 × 10−16 a,bVIII b
(Z)-3-hexenyl butyrate(1.234 ± 0.159) × 10−16 aVIII a
1.789 × 10−16 a,bVIII b
2-methyl-3-buten-2-ol(1.0 ± 0.03) × 10−17 a291 ± 1 II[237]
(8.3 ± 1.0) × 10−18 a293 ± 21IX[242]
(8.6 ± 2.9) × 10−18298 ± 20.974 ± 0.007X[224]
(8.6 ± 1.0) × 10−18 a298 ± 21XI[177]
1-octen-3-ol(5.00 ± 0.58) × 10−24 a2981XII[48]
Methyl salicylate~4 × 10−21 c298 ± 51XIII[164]
(3.33 ± 2.01) × 10−192981XIV[227]
a absolute rate constant; b calculated theoretically; c rough estimate due to low conversion of the reference reactant toluene.
Table 24. Experiments used to determine the rate constants listed in Table 23.
Table 24. Experiments used to determine the rate constants listed in Table 23.
ExptPhotoreactorDetectionAdditionReference ReactantsRef.
TypeVol. dm3Material
ICork
LISA
3910
977
Teflon
Pyrex
LP FTIR, O3
analyzer
no[201]
IIChamber3500–3700TeflonUV photometerCyclohexane ano[203,237,240]
III a O3 analyzerCyclohexane ano[238]
III bDFT method at the BH&HLYP/6-31+G(d,p) level of theory
IVChamber480PyrexLP FTIRNo b1,4-cyclohexadiene, isoprene[239]
VChamber775TeflonGC-MSFluorobenzene c no[92]
VIChamber6700
7900
TeflonGC-FIDCyclohexane a(E)-2-butene, propene, 2-methyl-2-butene[174]
VIIChamber6000PFALP FTIRCO a1,3,5-trimethylbenzene[160]
VIII aFlow tube0.012 O3 analyzer no[241]
VIII bDFT calculations with the BHandHLYP functional and the 6-311+G(d,p) basis sets at the BHandHLYP/6-311+G(d,p) level of theory.
IXReactor977PyrexLP FTIRCO ano[242]
XChamber480TeflonLP FTIRPropane aPropene, isobutene[224]
XILISA
CRAC
980
3910
Pyrex
FEP
LP FTIR
GC-MS
CO ano[177]
XIIChamber775
8000
TeflonNIR LDI MS, O3 analyzer no[48]
XIIIChamber50TeflonGC-FID Toluene[164]
XIVICARE-CHRS
chamber
7300TeflonHR PTR-TOF-MS 3,3,3-trifluoropropene[227]
a OH scavenger; b see comment in the text; c internal standard.
Table 25. Relative rate constants for gas-phase reactions of GLV with O3− Arrhenius parameters and values at selected temperature T (obtained by relative methods if not marked absolute).
Table 25. Relative rate constants for gas-phase reactions of GLV with O3− Arrhenius parameters and values at selected temperature T (obtained by relative methods if not marked absolute).
GLVAEA/RkTTT RangePExptRef.
Molecules cm−1 s−1KMolecules cm−1 s−1KKAtmTable 26
1-penten-3-ol(1.82 ± 2.08) × 10−16730 ± 348(1.61 ± 0.21) × 10−17298 ± 2273–3331I a[243]
(1.75 ± 0.25) × 10−17298 ± 2 I b
0.20 × 10−17298 I c
(Z)-2-penten-1-ol(2.32 ± 1.94) × 10−15902 ± 265(11.90 ± 1.40) × 10−17298 ± 2273–3331I a[243]
(9.07 ± 1.29) × 10−17298 ± 2 I b
1.50 × 10−17298 I c
(E)-2-penten-1-al(1.38 ± 0.73) × 10−16 a 1406 ± 163(1.24 ± 0.06) × 10−18 a298233–3730.8–1II[244]
(E)-3-hexen-1-ol(1.74 ± 1.65) × 10−151020 ± 300(5.97 ± 0.99) × 10−17298 ± 2273–3331I a[243]
(6.50 ± 0.95) × 10−17298 ± 2 I b
1.48 × 10−17298 I c
(E)-2-hexenal(1.79 ± 0.54) × 10−16 a 1457 ± 90(1.37 ± 0.03) × 10−18 a298233–3730.8–1II[244]
(Z)-2-hexenyl acetate(2.02 ± 0.65) × 10−15 c1306 ± 44 c2.50 × 10−17 b298200–3700.996III[196]
(E)-2-hexenyl
acetate
(7.21 ± 0.93) × 10−15 c1856 ± 33 c1.39 × 10−17 b
(Z)-3-hexenyl
acetate
(2.83 ± 0.39) × 10−15 c993 ± 35 c9.84 × 10−17 b
(E)-3-hexenyl
acetate
(1.68 ± 0.23) × 10−14 c925 ± 34 c7.37 × 10−16 b
Methyl salicylate2.97 × 10−12 b,d6185.62.9 × 10−21 d298278–3501IV[189]
a absolute rate constant; b practically did not depend on pressure in the range of 0.01–10,000 mmHg; c evaluated based on the reference data; d formation of primary ozonide.
Table 26. Experiments used to determine the rate constants listed in Table 25.
Table 26. Experiments used to determine the rate constants listed in Table 25.
ExpPhotoreactorDetectionAdd.Reference
Reactants
Ref.
TypeVol. dm3Material
I aFlow tube63PyrexLP FTIRCyclohexane a2,3-dimethyl-1,3- butadiene[243]
I bGC-MS (SPME)1-heptene
I cDFFT method with M06–2X functional and 6-311 þ G(d,p) triple split valence basis set
IIFlow tube63PyrexLP FTIRCyclohexane ano[244]
IIIDFT and modified RRKM[196]
IVDFT using B3LYP, M06-2X and MPW1K functionals with 6-311++G(d,p) basis set.[189]
a OH scavenger.
Table 27. Relative rate constants for gas-phase reactions of GLV with Cl at single temperatures.
Table 27. Relative rate constants for gas-phase reactions of GLV with Cl at single temperatures.
GLVK
cm3 Molecule−1 s−1
T
K
P
Atm
Expt
Table 28
Ref.
Pentan-1-ol(2.9 ± 0.2) × 10−102981I[210]
(2.51 ± 0.13) × 10−10298 ± 21II[212]
(2.57 ± 0.25) × 10−10295 ± 21III[213]
1-penten-3-ol(2.96 ± 1.22) × 10−10258 ± 11IV[205]
(2.37 ± 0.38) × 10−10262 ± 11 [205]
(2.40 ± 0.54) × 10−10273 ± 11 [205]
(2.35 ± 0.31) × 10−10298 ± 11 [205]
(2.78 ± 0.30) × 10−10313 ± 11 [205]
(2.57 ± 0.28) × 10−10333 ± 11 [205]
(Z)-2-penten-1-ol(2.99 ± 0.53) × 10−10296 ± 21V[245]
(3.00 ± 0.49) × 10−10298 ± 11IV[205]
(2.66 ± 0.47) × 10−10313 ± 11 [205]
(3.26 ± 0.48) × 10−10333 ± 11 [205]
(E)-2-pentenal(1.31 ± 0.19) × 10−102981VI[246]
1-penten-3-one(2.91 ± 1.10) × 10−102981VII[247]
(1.9 ± 0.4) × 10−10297–4001VIII[248]
(E)-2-hexen-1-ol(3.41 ± 0.65) × 10−10296 ± 21V[245]
(3.49 ± 0.82) × 10−10298 ± 31IX[249]
(Z)-3-hexen-1-ol(3.15 ± 0.58) × 10−10296 ± 21V[245]
(2.94 ± 0.72) × 10−10298 ± 31IX[249]
(E)-3-hexen-1-ol(3.05 ± 0.59) × 10−10296 ± 21V[245]
(3.42 ± 0.79) × 10−10298 ± 31IX[249]
(Z)-3-hepten-1-ol(3.80 ± 0.86) × 10−10298 31 [249]
Hexanal(2.88 ± 0.37) × 10−102981VI[246]
(3.23 ± 0.15) × 10−10298 ± 21X[220]
(E)-2-hexenal(1.92 ± 0.22) × 10−102981VI[246]
(Z)-3-hexenyl formate(2.45 ± 0.15) × 10−102981IV[222]
4.94 × 10−10 a2981 [222]
(E)-2-hexenylacetate(3.10 ± 1.13) × 10−10298 ± 1750 mmHgXI[186]
2-methyl-3-buten-2-ol(4.7 ± 1.0) × 10−10298 ± 20.974 ± 0.007XII[224]
(3.3 ± 0.4) × 10−10295 ± 10.921XIII[179]
1-octen-3-ol(4.03 ± 0.77) × 10−10296 ± 21V[245]
(Z)-3-octen-1-ol(4.13 ± 0.68) × 10−10298 ± 31IX[249]
Nonanal(4.82 ± 0.20) × 10−10298 ± 21X[220]
methyl salicylate(2.8 ± 0.44) × 10−12298 ± 51XIV[164]
(1.65 ± 0.3) × 10−122981XV[227]
a from SAR or LFER.
Table 28. Experiments used to determine the rate constants listed in Table 27.
Table 28. Experiments used to determine the rate constants listed in Table 27.
ExptPhotoreactorDetectionCl SourceAdd.Reference ReactantsRef.
TypeVol. dm3Material
IBag140TeflonGC-FIDCl2 Pentane
1,3-dioxolane
Cyclohexane
[210]
IICylinder50TeflonGC-FIDCl2, COCl2 Cycloxexane[212]
IIIBag100TeflonGC-FIDCl2 Propane;
hexane
[213]
IVChamber400TeflonGC-FIDCCl3COCl Octane, propene,
cyclohexane
[205,222]
VChamber600TeflonIon flow tupe MSCl2 Tetrahydrofuran,
propan-1-ol, octane
[245]
VIChamber200TeflonGC-FIDCl2 Ethene; propene;
1-butene
[246]
VIIBag80TeflonGC-FIDClC(O)C(O)Cl Chloroethene
1,1-dichloroethene
acrylonitrile
[247]
VIIIBulb0.5PyrexGC-FID; GC-MS; FTIRCl2 Ethane[248]
IXChamber480PyrexLP FTIRCl2 1-butene, isobutene[249]
XCylinder108OPyrexLP FTIRCl2 (E)-2-butene;
Methy lvinyl ketone
[220]
XIBag80TeflonGC-FIDClC(O)C(O)Cl (n-butyl methacrylate
n-butyl acrylate
[186]
XIIChamber480TeflonLP-FTIRCl2 Ethene[224]
XIIIChamber47SteelLP-FTIRCl2 Ethane, cyclohexane[179]
XIVChamber50TeflonGC-FIDCl2 Acetone[164]
XVICARE-CNRS chamber7300TeflonHR PTR-TOF-MSCl2 Acetone[227]
Table 29. Relative and absolute (a) rate constants for gas-phase reactions of GLV with Cl atoms—Arrhenius parameters and values at selected temperature T.
Table 29. Relative and absolute (a) rate constants for gas-phase reactions of GLV with Cl atoms—Arrhenius parameters and values at selected temperature T.
GLVAEA/RkTTT RangePExptRef.
cm3 Molecule−1 s−1Kcm3 Molecule−1 s−1KKAtmTable 30
hexanal(7.91 ± 0.66) × 10−11 a−(349 ± 51)(2.56 ± 0.24) × 10−11 a298265–3811I[250]
2-methyl-3-butene-2-ol(2.83 ± 2.50) × 10−14−(2670 ± 249)(2.13 ± 0.19) × 10−10298256–2981II[251]
methyl salicylate1.2703 × 10−8 a,b,c1438.41.01 × 10−10 a,b,c,d298278–3501III[209]
a absolute rate constants; b determined theoretically; c calculated here from the reference data; d 100 times higher than experimental values from Table 23 [209].
Table 30. Experiments used to determine the rate constants listed in Table 29.
Table 30. Experiments used to determine the rate constants listed in Table 29.
ExptPhotoreactorDetectionCl SourceAdd.Reference
Reactants
Ref.
TypeVol. dm3Material
ICell (PLP)0.2PyrexRFCl2NoNo[250]
IIChamber400TeflonGC-FIDCCl3COCl Cycloxexane, methanol,
propene,
1-butene
[251]
IIIDFT methods at B3LYP and M06-2X levels of theory with 6-311++G(d,p) basis set[209]
Table 31. GLV photolysis rate constants j, absolute, and relative to the photolysis rate constant of NO2, determined at given effective quantum yields Φ and zenith angles θ.
Table 31. GLV photolysis rate constants j, absolute, and relative to the photolysis rate constant of NO2, determined at given effective quantum yields Φ and zenith angles θ.
GLVj
s−1
j/j(NO2)Φθ
°
T
K
Expt
Table 32
Ref.
1-penten-3-ol(1.61–2.36) × 10−6 a 1 298I[57]
(E)-2-pentenal3.06 × 10−4 120298II[162]
2.87 × 10−4 30
2.58 × 10−4 40
2.18 × 10−4 50
1.63 × 10−4 60
9.73 × 10−5 70
1-penten-3-one(0.36−1.39) × 10−5 a 1 298I[57]
n-hexanal1.7 × 10−50.2 × 10−20.28 286–294III[161]
1.14 × 10−5 0.28 298I[159]
(1.81 ± 0.10) × 10−5 0.31 ± 0.0217298IV[220]
(Z)-3-hexenal2.1 × 10−50.4 × 10−20.34 286–294III[161]
(2.61 ± 0.08) × 10−5(4.7 ± 0.4) × 10−30.25 ± 0.06 298V[160]
(E)-2-hexenal1.0 × 10−41.8 × 10−20.36 286–294III[161]
3.80 × 10−4 120298II[162]
4.05 × 10−4 30
3.80 × 10−4 40
3.42 × 10−4 50
2.89 × 10−4 60
2.17 × 10−4 70
(Z)-3-hexen-1-ol(1.61−2.36) × 10−5 a 1 298I[57]
Methyl salicylate(2.82 ± 0.26) × 10−5 b
Insignificant c
298VI[227]
a upper limits, 0–10 km above the Earth’s surface; b at 254 nm; c at 365 nm and > 300 nm.
Table 32. Experiments used to determine the rate constants listed in Table 31.
Table 32. Experiments used to determine the rate constants listed in Table 31.
ExptPhotoreactorDetectionAdditionsRef.
TypeVol. dm3Material
ICalculation conditions: clear-sky, July noon, Ciudad Real (Spain). The actinic flux was based on the TUV model a.[57,159]
IIThe actinic flux was based on the TUV model a.[162]
IIIEUPHORE2 × 105PTFELP-FTIR, GC-FID, GC-MS [161]
IVEUPHORE2 × 105PTFELP-FTIR [220]
VChamber6000PFALP-FTIRCO b, isoprene c[160]
VIICARE-CNRS7300FEPHR PTR-TOF-MS [227]
a Tropospheric Ultraviolet Model [252]; b OH scavenger; c OH tracer.
Table 33. Relative rate constants for aqueous-phase reactions of GLV with OH radicals—Arrhenius parameters and values at 298 K and pH = 5.4, and yields of SOA products.
Table 33. Relative rate constants for aqueous-phase reactions of GLV with OH radicals—Arrhenius parameters and values at 298 K and pH = 5.4, and yields of SOA products.
GLVAEAT RangepHk298 K aSOARef.
1011 M−1 s−1kJ mol−1K 108 M−1 s−1%
1-penten-3-ol10.4 ± 0.313 ± 2278–31876.3 ± 0.1-[163]
1-penten-3-one 7.2 b-[215]
(Z)-2-hexen-1-ol2.6± 0.19 ± 2273–31876.7 ± 0.3-[163]
(Z)-3-hexen-1-ol8.1 ± 0.912 ± 0.3278–2985.45.1 ± 0.852[22]
(E)-2-hexenal5.2 ± 0.112 ± 1273–31874.8 ± 0.3-[163]
(Z)-3-hexenyl acetate190 ± 4017 ± 2278–2985.48.7 ± 1.18[22]
methyl salicylate58 ± 4014 ± 1278–2985.47.8 ± 0.587
methyl jasmonate58 ± 915 ± 2278–2985.46.8 ± 0.867
2-methyl-3-butene-2-ol23 ± 513 ± 2278–2985.47.5 ± 1.420
a experimental values; b estimated from SAR or LFER.
Table 34. Relative rate constants for aqueous-phase reactions of GLV with SO4 and NO3 radicals at pH = 7—Arrhenius parameters (278–318 K) and values at 298 K [163].
Table 34. Relative rate constants for aqueous-phase reactions of GLV with SO4 and NO3 radicals at pH = 7—Arrhenius parameters (278–318 K) and values at 298 K [163].
GLVXAEAk298 K
109 M−1 s−1kJ mol−1108 M−1 s−1
1-penten-3-olSO47.9 ± 0.15 ± 19.4 ± 1.0
NO3150 ± 1017 ± 21.5 ± 0.2
(Z)-2-hexen-1-olSO4110 ± 1010 ± 228.3 ± 3
NO338 ± 19 ± 18.4 ± 2.3
(E)-2-hexenalSO42.9 ± 0.14 ± 14,8 ± 0.2
NO331 ± 117 ± 20.28 ± 0.07
Table 35. Relative rate constants for aqueous-phase reactions of GLV with singlet molecular oxygen 1O2 * and 3C * triplet state (3DMB * and 3MAP *) at pH = 5.4, and yields of some SOA products [265].
Table 35. Relative rate constants for aqueous-phase reactions of GLV with singlet molecular oxygen 1O2 * and 3C * triplet state (3DMB * and 3MAP *) at pH = 5.4, and yields of some SOA products [265].
GLV1O2 *3DMB *3MAP *
AEAk298 KkSOA k
M−1 s−1kJ mol−1106 M−1 s−1106 M−1 s−1%106 M−1 s−1
pH = 2.1
298 K
pH = 5.1
278–298 K
pH = 5.1
(Z)-3-hexen-1-ol6.1 × 102082 ± 7.42.5 ± 0.30.33 ± 0.040.24 ± 0.1-1.2 ± 0.8
(Z)-3-hexenyl acetate2.2 × 101550 ± 7.23.9 ± 0.814 ± 214 ± 7387.3 ± 2
methyl salicylate--≤ 0.129 ± 212 ± 4808.0 ± 0.8
methyl jasmonate3.6 × 102396 ± 4.86.0 ± 0.73.6 ± 0.54.2 ± 3841.2 ± 0.5
2-methyl-3-butene-2-ol6.7 × 10922 ± 1.77.5 ± 1.40.28 ± 0.10.13 ± 0.07-0.55 ± 0.2
Table 36. Pseudo-first order rate constant for the aqueous-phase reaction of MeJa with OH radicals in bulk solutions and aqueous films.
Table 36. Pseudo-first order rate constant for the aqueous-phase reaction of MeJa with OH radicals in bulk solutions and aqueous films.
Film Thickness, µm∞ (Bulk)193.177.238.6
k1st, 10−4 min−12.83 ± 0.029.62 ± 0.4311.0 ± 0.512.7 ± 0.6
Table 37. Products identified in smog chamber oxidation of (Z)-3-hexen-1-ol by OH and O3.
Table 37. Products identified in smog chamber oxidation of (Z)-3-hexen-1-ol by OH and O3.
NameOxidantMWFormulaStructurePhaseRef.
acetaldehydeO344C2H4O Atmosphere 12 01655 i036gas[92]
2-propenal56C3H4O Atmosphere 12 01655 i037
propanal58C3H6O Atmosphere 12 01655 i038
acetic acid60C2H4O2 Atmosphere 12 01655 i039
2-propenoic acid72C3H4O2 Atmosphere 12 01655 i040
3-hydroxypropanal74C3H6O2 Atmosphere 12 01655 i041PM[198]
propionic acid74C3H6O2 Atmosphere 12 01655 i042Gas, PM[92,198]
2-hydroxyacetic acid76C2H4O3 Atmosphere 12 01655 i043PM[198]
3-hydroxy-2- oxopropanal88C3H4O3 Atmosphere 12 01655 i044
2,3-di-hydroxypropanal90C3H6O3 Atmosphere 12 01655 i045
2-hydro-peroxypropanal90C3H6O3 Atmosphere 12 01655 i046
3-hydroxy-2-
oxopropanoic acid
104C3H4O4 Atmosphere 12 01655 i047
2-hydroperoxy-3-
hydroxypropanal
106C3H6O4 Atmosphere 12 01655 i048
2-ethyl-1,3-dioxan-4-olOH, O3132C6H12O3 Atmosphere 12 01655 i049PM[65,198]
3-(2-hydroxy ethoxy) propanoic acid134C5H10O4 Atmosphere 12 01655 i050[65]
2-(2-hydroxyethyl)-
1,3-dioxan-4-ol
148C6H12O4 Atmosphere 12 01655 i051[65,198]
2-(1,3-dioxin-2-
yl) ethylformate
O3158C7H10O4 Atmosphere 12 01655 i052PM[198]
3-(3–hydroxy propanoyloxy)
propanoic acid
OH, O3162C6H10O5 Atmosphere 12 01655 i053PM[65]
2-(4-hydroxy-1,3-
dioxan-2-yl)ethyl formate
O3176C7H12O5 Atmosphere 12 01655 i054PM[198]
2-(1,3-dioxin-2-yl)
ethyl propionate
186C9H14O4 Atmosphere 12 01655 i055
1-((2-ethyl-1,3-dioxan-
4-yl)oxy)propan-1-ol
190C9H18O4 Atmosphere 12 01655 i056
2-(2-((3-hydroxyprop-
1-en-1-yl)oxy)ethyl)-
1,3-dioxan-4-ol
204C9H16O5 Atmosphere 12 01655 i057
2-(4-hydroxy-1,3-
dioxan-2-yl)ethyl
propionate
204C9H16O5 Atmosphere 12 01655 i058
3-(3-(2-
hydroxyethoxy)
propanoyloxy) propanoic acid
OH, O3206C8H14O6 Atmosphere 12 01655 i059PM[65]
3-(formyloxy)-3-(2-
oxoethoxy) propyl propionate
O3218C9H14O6 Atmosphere 12 01655 i060PM[198]
3-(3-(2-hydroxy
ethoxy)-3-
oxopropanoyloxy)
propanoic acid
OH, O3220C8H12O7 Atmosphere 12 01655 i061PM[65]
3-(3-(2-hydroxyethoxy)-3-hydroxypropanoyloxy) propanoic acid222C8H14O7 Atmosphere 12 01655 i062
1-(2-(4-hydroxyl-1,3-dioxan-2-yl)-ethoxy)-propane-1,3-diol222C9H18O6 Atmosphere 12 01655 i063
1,3-dihydrox-3- (2-hydroxyethoxy)
propyl- 3- hydroxy propanoate
224C8H16O7 Atmosphere 12 01655 i064
3-( 3-(3-hydroxy propanoyloxy)
Propanoyloxy propanoic acid
234C9H14O7 Atmosphere 12 01655 i065
3-(carboxyoxy)-3-(2-
oxoethoxy) propyl
propionate
O3234C9H14O7 Atmosphere 12 01655 i066PM[198]
1-(1-((2-ethyl-1,3-
dioxan-4-yl)oxy)propoxy)
propan-1-ol
248C12H24O5 Atmosphere 12 01655 i067
3-(3-(3-(3-hydroxy propanoyloxy)
propanoyloxy)
propanoyloxypropanoic acid
OH, O3306C12H18O9 Atmosphere 12 01655 i068PM[65]
Table 38. Products identified in smog chamber ozonolysis of (Z)-3-hexenyl acetate.
Table 38. Products identified in smog chamber ozonolysis of (Z)-3-hexenyl acetate.
NameMWFormulaStructurePhaseRef.
2-propenal56C3H4O Atmosphere 12 01655 i069Gas[92]
Propanal58C3H6O Atmosphere 12 01655 i070
Acetic acid60C2H4O2 Atmosphere 12 01655 i071
2-propenoic acid72C3H4O2 Atmosphere 12 01655 i072
Propionic acid74C3H6O2 Atmosphere 12 01655 i073Gas, PM[92,198]
2-hydroperoxy propanal90C3H6O3 Atmosphere 12 01655 i074PM[198]
3-oxo-propyl acetate116C5H8O3 Atmosphere 12 01655 i075PM[272]
2-acetoxyacetic acid118C4H6O4 Atmosphere 12 01655 i076PM[198]
2,3-dioxopropyl acetate130C5H6O4 Atmosphere 12 01655 i077
2-hydroxy-3-oxopropyl
acetate
132C5H8O4 Atmosphere 12 01655 i078
3-acetoxy-propanoic acid132C5H8O4 Atmosphere 12 01655 i079PM[198,272]
3-acetoxy-2-oxopropanoic acid146C5H6O5 Atmosphere 12 01655 i080PM[198]
2-hydroperoxy-3-
oxopropyl acetate
146C5H8O5 Atmosphere 12 01655 i081
3-acetoxypropane peroxoic acid148C5H8O5 Atmosphere 12 01655 i082PM[272]
3,4-dioxohexyl acetate172C8H12O4 Atmosphere 12 01655 i083PM[198]
2-hydroxyethyl 3-
acetoxypropanoate
176C7H12O5 Atmosphere 12 01655 i084PM[272]
2-(3-oxopropyl)ethyl 3-
acetoxy propanoate
232C10H16O6 Atmosphere 12 01655 i085PM[272]
3-acetoxypropanoyl 3-
acetoxypropanoate
246C10H14O7 Atmosphere 12 01655 i086PM[198]
2-(2-(3-acetoxy
propanoyloxy)ethoxy)
propanoic acid
248C10H16O7 Atmosphere 12 01655 i087PM[272]
5-acetoxy-3-oxopentyl-3-acetoxypropanoate274C12H18O7 Atmosphere 12 01655 i088
2-(3-
acetoxypropanoyloxy)ethyl 3-acetoxypropanoate
290C12H18O8 Atmosphere 12 01655 i089
[3-[2-(3- acetoxypropoxy)ethoxy]-3-oxo-propyl] 3-acetoxypropanoate348C15H24O9 Atmosphere 12 01655 i090
Table 39. Products identified in SOA from smog-chamber and aqueous-phase reactions of 2-methyl-3-buten-2-ol and ambient SOA samples (except organosulfates listed in Table 40).
Table 39. Products identified in SOA from smog-chamber and aqueous-phase reactions of 2-methyl-3-buten-2-ol and ambient SOA samples (except organosulfates listed in Table 40).
MWNameFormulaStructurePhaseRef.Ambient Aerosol
30formaldehydeCH2O Atmosphere 12 01655 i091Gas[273,274]
32methanolCH4O Atmosphere 12 01655 i092Aqu[268] b
44acetaldehydeC2H4O Atmosphere 12 01655 i093Gas[273]
46formic acidCH2O2 Atmosphere 12 01655 i094Aqu[268] b
48formaldehyde hydrated
(methanediol)
CH4O2 Atmosphere 12 01655 i095
58acetoneC3H6O Atmosphere 12 01655 i096Gas, aqu[268,273,274] b
58glyoxalC2H2O2 Atmosphere 12 01655 i097Gas[187,273]
60glycolaldehydeC2H4O2 Atmosphere 12 01655 i098
61acetic acidC2H4O2 Atmosphere 12 01655 i099Aqu[268] b
72methylglyoxalC3H4O2 Atmosphere 12 01655 i100Gas[273]
74glyoxylic acidC2H2O3 Atmosphere 12 01655 i101
76glycolic acidC2H4O3 Atmosphere 12 01655 i102Aqu[268] b
78glycolaldehyde hydratedC2H6O3 Atmosphere 12 01655 i103
861,3-butanedioneC4H6O2 Atmosphere 12 01655 i104Gas[273]
862-oxopropanedialC3H2O3 Atmosphere 12 01655 i105
882-hydroxy-2-methylpropanalC4H8O2 Atmosphere 12 01655 i106Gas[187,273]
882-hydroxypropaneialC3H4O3 Atmosphere 12 01655 i107Gas[273]
1002,3-dioxobutanalC4H4O3 Atmosphere 12 01655 i108
1142,3-dioxobutane-1,4-dialC4H2O4 Atmosphere 12 01655 i109
1162-oxovaleric acidC5H8O3 Atmosphere 12 01655 i110
1181,3-dihydroxy-3-methyl-
butan-2-one
C5H10O3 Atmosphere 12 01655 i111Gas, PM
1182,3-dihydroxy-3-
methylbutanal
C5H10O3 Atmosphere 12 01655 i112Gas
1202,3-dihydroxyisopentanolC5H12O3 Atmosphere 12 01655 i113PM, aqu[266] a, [268] b, [273]PM2.5 [273,275]
1322,3-dihydroxy-2-methylbutane dialdehydeC5H8O4 Atmosphere 12 01655 i114Gas, PM[273]
1342,3-dihydroxy-3-methyl-
butanoic acid
C5H10O4 Atmosphere 12 01655 i115PM
1342-hydroxy-2-
methylpropenedioic acid
C4H6O5 Atmosphere 12 01655 i116
1362-methylerythritolC5H12O4 Atmosphere 12 01655 i117PM2.5 [273,275]
1362-methylthreitolC5H12O4 Atmosphere 12 01655 i118PM2.5 [273,275]
1642,3-dihydroxy-2-
methylsuccinic acid c
C5H8O6 Atmosphere 12 01655 i119PM2.5 [276]
a aqueous phase reactions of methylbutenol epoxides (Section 5); b aqueous-phase addition of sulfate radical anions (Section 5); c and isomers.
Table 40. Organosulfates identified in smog chamber and aqueous-phase experiments with GLV, eventually confirmed in samples of ambient aerosols.
Table 40. Organosulfates identified in smog chamber and aqueous-phase experiments with GLV, eventually confirmed in samples of ambient aerosols.
ProductParent
Compound
Formation ProcessRef.Ambient Aerosol
MWNameFormulaStructure
140Glycolaldehyde sulfateC2H4O5S Atmosphere 12 01655 i120MBOSO4 addition[268] bPM2.5 [277]
PM10 [278]
154Hydroxyacetone sulfateC3H6O5S Atmosphere 12 01655 i121(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM1 [280,281], PM2.5 [277,279,282,283,284,285], PM10 [278]
156 Glycolic acid
sulfate
C2H4O6S Atmosphere 12 01655 i122MBOSO4 addition[268] bPM1 [280,281,286],
PM2.5 [277,282,283,284,285,287,288]
158Hydrated glycol sulfateC2H6O6S Atmosphere 12 01655 i123MBOSO4 addition[268] b
170Lactic acid sulfateC3H6O6S Atmosphere 12 01655 i124(E)-2-pentenalOzonolysis[289]PM1 [280,286], PM2.5 [277,282,283,284,287,289,290,291,292]
1701-sulfooxy-2-
hydroxybutane
C4H10O5S Atmosphere 12 01655 i125(E)-2-pentenalOzonolysis[289]PM2.5 [289]
1702-Sulfoxy-3-
hydroxy-propanal
C3H6O6S Atmosphere 12 01655 i126(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM2.5 [279]
1863-sulfoxy-2-
hydroxy-propanoic acid
C3H6O7S Atmosphere 12 01655 i127(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM2.5 [279]
1983-hydroxy-3-
methyl-butan-2-one sulfate
C5H10O6S Atmosphere 12 01655 i128MBOSO4 addition[268] b
1984-sulfoxy-1-
hydroxy-3-
m3thyl-butan-2-one
C5H10O6S Atmosphere 12 01655 i129MBOSO4 addition[268] bPM2.5 [277],
PM2.5 [288]
2002,3-dihydroxy-3-methyl-butane
sulfate
C5H12O6S Atmosphere 12 01655 i130MBOOH photolysis
SO4 addition
[266] a, [268] b,[293]PM1 [280,281] ?, PM2.5 [277,283,293], [288] ?
210(Z)-5-sulfoxy-
hex-3-enoic acid
C6H10O6S Atmosphere 12 01655 i131(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM1 [280,281] ?, PM2.5 [279],
PM10 [278] ?
2126-(sulfoxy)
hexanoic acid
C6H12O6S Atmosphere 12 01655 i132(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM1 [280,281] ?, PM2.5 [279]
2143-sulfoxy-2-
hydroxypentanoic acid
C5H10O7S Atmosphere 12 01655 i133(E)-2-pentenal (E)-2-pentenoic acidOzonolysis, SO4 (aqu)[289]PM1 [280] ?,
PM2.5 [289]
2142-sulfoxy-3-
hydroxypentanoic acid
C5H10O7S Atmosphere 12 01655 i134(E)-2-pentenalOzonolysis[289]PM1, [280,281] ?,
PM2.5 [289]
226(E)-5-sulfoxy-4-
hydroxy-hex-2-enoic acid
C6H10O7S Atmosphere 12 01655 i135(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM2.5 [279]
226 C6H10O7S Atmosphere 12 01655 i136(Z)-3-hexenalOzonolysis[294]PM2.5 [294]
2303-sulfooxy-2,4-
dihydroxypentanoic acid
C5H10O8S Atmosphere 12 01655 i137(E)-2-pentenal, (Z)-3-hexenal (Z)-2-hexenalOzonolysis ozonolysis[289]PM2.5 [289]
270 C9H18O7S-(Z)-3-hexen-1-olOH
photolysis, ozonolysis
[279]PM2.5 [279]
a aqueous phase reactions of methylbutenol epoxides (Section 5); b aqueous-phase addition of sulfate radical anions (Section 5); ? unresolved structure.
Table 41. Atmospheric lifetimes of GLV at 298 K due to gas-phase reactions with OH (2 × 106 molecules cm−3), NO3, (5 × 108 molecules cm−3), O3 (7 × 1011 molecules cm−3), and Cl (1 × 104 molecules cm−3).
Table 41. Atmospheric lifetimes of GLV at 298 K due to gas-phase reactions with OH (2 × 106 molecules cm−3), NO3, (5 × 108 molecules cm−3), O3 (7 × 1011 molecules cm−3), and Cl (1 × 104 molecules cm−3).
GLVOHNO3O3ClPhotolysis
τ
h
Ref.τ
h
Ref.τ
h
Ref.τ
h
Ref.τ
h
Zenith
Angle, °
Ref.
Pentan-1-ol24 b[184]
1-penten-3-ol2.1[185]40[192] a23.5[201]118[205] a174–120 c [57]
2.83[57] b 22.3[237] a
1.95[228] a 24.3[243] b
(Z)-2-penten-1-ol1.3[185]3.6[192] a3.7[201]93[245]
2.5[237] a92.6[205] a
2.1[243] b
(E)-2-penten-1-ol2.05[57] a
(E)-2-pentenal5.9[85] a28.8[230] a309[244] b212[246] b0.7326[162]
3.23[173] a
1-penten-3-one1.18[57] b16.4[215] a29.2[201] 77–20 ac [57]
hexanal4.4[85] a32[232] a 97[246] b24.4–13.4 ac [159]
5.0[159] a
(Z)-2-hexen-1-ol1.25[217] b3.6[231]5.3[238]
4.27[171] b1.5[230]
(E)-2-hexen-1-ol1.4[176]4.3[231]6.7[175]81[245]
4.04[171] b 2.4[238]81[249]
2.26[57] a
(Z)-3-hexen-1-ol1.3[174,228] a2.1[231]3.8[237,240] a91[245]17.4–11.8 ac [57]
1.0[176]2.0[174]2.8[243] b95[249]
4.78[57] b 6.6[175]
5.05[171] b 6.2[174]
7.2[238]
(E)-3-hexen-1-ol1.2[176]1.3[231]6.9[175]86[245]
1.25[175] 6.4[238]81[249]
1.75[217] b 6.6[243] a
4.55[171] b
(Z)-4-hexen-1-ol3.92[171] b1.9[231]5.6[238]
(E)-4-hexen-1-ol3.57[171] b 3.8[238]
(Z)-2-hexenal 40.8[230] a
(E)-2-hexenal3.12[174] b81.2[174]192[174]145[246] b0.4826[162]
2.97[159] a41[230]274[244] b 2.7 [161]
10[235]
125[193]
(Z)-3-hexenal 11.5[160] a 13.3 [161]
5.630[160]
(Z)-3-hepten-1-ol1.09[175] 73[249]
1-octen-3-ol 69[245]
(Z)-3-octen-1-ol0.94[175] 67[249]
2-methyl-3-buten-2-ol3.56[224] a26.5[234] b39.9[237] a130[251] a>5 yrs [165]
2.01[179] a46[236]47.8[242] a59.1[224] a
2.17[165] a 46.1[224] a84.2[179] a
2.5[177]
(E)-2-hexenylacetate2[186] 9.0[186]
(Z)-3-hexenyl formate3[222] b 9.8[241]113.5[222]
(Z)-3-hexenyl acetate1.76[174] b2.3[174]7.3[174]
6.9[241]
4.0[196] a
(Z)-3-hexenyl propionate 5.2[241]
(Z)-3-hexenyl butyrate 3.2[241]
Methyl salicylate70.88[190] 99192 11.3 yrs[164] b15378 1.64 yr[164] b
15.6 yrs[189] a275 b[209]
a calculated using the rate constant from the reference; b recalculated from the reference values to another concentration; c at 0–10 km above Earth’s surface.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sarang, K.; Rudziński, K.J.; Szmigielski, R. Green Leaf Volatiles in the Atmosphere—Properties, Transformation, and Significance. Atmosphere 2021, 12, 1655. https://doi.org/10.3390/atmos12121655

AMA Style

Sarang K, Rudziński KJ, Szmigielski R. Green Leaf Volatiles in the Atmosphere—Properties, Transformation, and Significance. Atmosphere. 2021; 12(12):1655. https://doi.org/10.3390/atmos12121655

Chicago/Turabian Style

Sarang, Kumar, Krzysztof J. Rudziński, and Rafał Szmigielski. 2021. "Green Leaf Volatiles in the Atmosphere—Properties, Transformation, and Significance" Atmosphere 12, no. 12: 1655. https://doi.org/10.3390/atmos12121655

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