Abstract

Methyl-butenols are a category of oxygenated biogenic volatile organic compounds emitted by plants as part of their natural metabolic processes. This study examines the gas-phase reactions of ozone (O3) with five methyl-butenols (2-methyl-3-buten-2-ol, 3-methyl-2-buten-1-ol, 3-methyl-3-buten-1-ol, 2-methyl-3-buten-1-ol, and 3-methyl-3-buten-2-ol) under atmospheric conditions at a temperature of (298 ± 2) K and pressure of (1000 ± 10) mbar. The experimental values for the gas-phase reaction rate coefficients obtained in this study, by using the relative rate method, are as follows (in cm3 molecule–1 s–1): k(3-methyl-2-buten-1-ol + O3) = (311 ± 20) × 10–18, k(2-methyl-3-buten-2-ol + O3) = (9.55 ± 1.04) × 10–18, k(3-methyl-3-buten-1-ol + O3) = (7.29 ± 0.46) × 10–18, k(2-methyl-3-buten-1-ol + O3) = (4.25 ± 0.29) × 10–18, and k(3-methyl-3-buten-2-ol + O3) = (62.9 ± 6.8) × 10–18. The results are discussed in detail, with particular emphasis on the degree and type of substitutions of the double bond. The determined rate coefficient values are also compared to the available literature data and with estimates of the structure–activity relationship. Additionally, the atmospheric implications toward the tropospheric lifetime and photochemical ozone generation potential for the investigated compounds are provided, which highlight the atmospheric impact of methyl-butenol decomposition into the lower atmosphere.
1. Introduction
Unsaturated oxygenated hydrocarbons in the form of alcohols, carbonyls, ethers, and esters are continuously emitted into the atmosphere.1 Plants release unsaturated oxygenated volatile organic compounds (VOCs) in correspondence with their physiological status associated with various development stages or as a response to various stressors.2 The oxidation processes initiated by radicals in the form of hydroxyl (OH) and nitrate (NO3), or by ozone (O3), represent the dominant sink mechanism of unsaturated oxygenated organic volatiles.3 Among the air pollutants, tropospheric ozone, formed by a complex reaction involving VOCs and nitrogen oxides (NOx), is considered the most phytotoxic with adverse effects on plants and with an important role in the chemical degradation of unsaturated VOCs. These chemical interactions result in the production of photochemically active species4 and secondary organic aerosols (SOA), which have been demonstrated to affect human health and influence the air quality and climate.5−7 As showed by O’Dwyer et al.,8 these compounds and their reaction products often manifest as haze over forests.
Methyl-butenols (MBO) are five-carbon unsaturated oxygenated alcohols produced enzymatically through the methylerythritol 4-phosphate (MEP) pathway in chloroplasts of pine forest leaves during the day, along with isoprene.9,10 Significant quantities of 2-methyl-3-butene-2-ol (MBO232), 3-methyl-2-butene-1-ol (MBO321), 3-methyl-3-butene-1-ol (MBO331), 2-methyl-3-butene-1-ol (MBO231), and 3-methyl-3-butene-2-ol (MBO332) were identified in the border layer near pine forests11,12 and flowering rye fields.13 MBO232 was also identified as a product of fungal symbionts of the spruce bark beetle in deciduous forests.14,15 While the gas-phase OH radical-initiated reactions of unsaturated alcohols are relatively well-known,16−19 reactions initiated by O3 molecules, NO3 radicals, or Cl atoms have received less attention. However, the last referred reactions represent important processes in terms of atmospheric chemistry, air quality, and environmental impact.20−23 Reactions between ozone and unsaturated alcohols are significant precursors for SOA and pollutant formation with effects on the ecosystems and photochemical smog formation.
The ozonolysis is known to produce gas-phase carbonyls and acids with a role in highly oxygenated molecule (HOM) formation,24 with contribution to acid deposition.25 The importance of the ozone reaction with unsaturated species, including alcohols, is also given by a significant contribution to the OH radical formation, which peaks during the night time (contributions up to 64% of the OH radical budget).26
The gas-phase reaction rate coefficient of ozone with MBO232 at atmospheric pressure and a temperature of (291 ± 1) K was initially determined by Grosjean and Grosjean27 in a 3.5 m3 collapsible Teflon chamber. Pseudo-first-order decay of ozone (O3), with cyclohexane used as an OH radical scavenger in 400:1 mixing ratio with MBO232, was employed for the experimental investigations. The reported value was (10.0 ± 0.3) × 10–18 cm3 molecule–1 s–1. Fantechi et al.28 conducted kinetic studies in a 480 L evacuable photoreactor with FTIR spectroscopy used to monitor the relative ozone-driven decay of MBO232. In the presence of propane as an OH radical scavenger, the authors reported a value of (8.6 ± 2.9) × 10–18 cm3 molecule–1 s–1 for the reaction of MBO232 with propene and isobutene as reference compounds for the relative kinetic study. Klawatsch-Carrasco et al.29 conducted experiments in a 977 L Pyrex evacuable photoreactor using the absolute rate method and the Fourier-transform infrared (FTIR) technique to measure both ozone and MBO232. In the presence of carbon monoxide (CO) used as an OH radical scavenger, a reaction rate coefficient of (8.3 ± 1.0) × 10–18 cm3 molecule–1 s–1 was found at (293 ± 3) K and 1 atm synthetic air. The rate coefficient for the reaction between MBO321 and ozone under atmospheric conditions was investigated by Grosjean and Grosjean.30 The researchers employed experimental techniques similar to those previously described and reported a value of (439 ± 37) × 10–18 cm3 molecule–1 s–1. With regard to the reaction of MBO331 and MBO231 with ozone, Gai et al.31 determined rate coefficient values of (6.80 ± 1.29) × 10–18 cm3 molecule–1 s–1 and (3.74 ± 0.62) × 10–18 cm3 molecule–1 s–1, respectively, using a 170 L collapsible Teflon bag and an ozone analyzer to monitor the pseudo-first-order decay of O3 molecules accompanied by cyclohexane as an OH radical scavenger.
The primary objectives of the present study are to provide the experimental reaction rate coefficients of five selected methyl-butenols with ozone (1–5) and to discuss their reactivity, emphasizing contributions of the structural features of homologous compounds, intramolecular occurring effects, structure–activity relationship (SAR) methodologies, and atmospheric implications. Additionally, another experimental approach that helps correct secondary chemical processes driven by the OH radicals formed during ozonolysis is also discussed. To the best of our knowledge, this study reports for the first time on the reaction rate coefficient of ozone and MBO332 under atmospherically relevant conditions.
| 1 |
| 2 |
| 3 |
| 4 |
| 5 |
Photochemical ozone creation potential (POCP) is a useful tool to quantify the relative abilities of different VOCs to generate ground-level ozone and to assess the impact of volatiles in the lower troposphere.32 The present study also provides the POCPE values for the five selected MBOs and addresses an extensive perspective regarding the reactivity and atmospheric impact of these atmospheric reactions.
2. Experimental Section
The reaction rate coefficients corresponding to the ozonolysis reaction of selected MBOs were experimentally determined under atmospheric conditions, at a temperature of (298 ± 2) K and an air pressure of (1000 ± 10) mbar, using the facilities provided by the 760 L Environmental Simulation Chamber made of Quartz from the “Alexandru Ioan Cuza” University of Iasi (ESC-QUAIC), Romania. Further details regarding the employed infrastructure and experimental setup can be found elsewhere.33 The relative rate method was employed throughout the experimental procedure, using at least two kinetic reference compounds and two different methods to handle the possible undesirable reactions of OH radicals with MBOs and reference compounds, radicals produced during the ozonolysis. Chemical transfer into the reaction vessel was achieved by injecting weighed amounts of organics through a heated inlet port, assisted by a stream of carrier gas. Two Teflon blade fans were employed to ensure the homogeneity of the reaction mixture. The chemical composition of the gas mixture was continuously monitored during the investigation by a Bruker Vertex 80 FT-IR spectrometer connected to a white-type multireflection mirror system installed inside the reactor and ensuring a total optical path of (492 ± 1) m. Ozone was produced by passing a flow of oxygen over a Hg lamp (λ = 184.9 nm) and transferred into the reaction vessel to initiate the oxidation processes. 1,3,5-Trimethylbenzene was added as an OH radical tracer compound to enable estimation of the OH concentration and to correct for the decay of MBOs and reference compounds due to the reaction with OH radicals. Alternatively, a high quantity of carbon monoxide or 1,3,5-trimethylbenzene was added to the gas mixture to scavenge over 90% of the OH radicals, thus enabling the experiments involving MBO321, MBO331, and MBO231 to be conducted.
The initial gas-phase concentrations (molecules cm–3) of the MBOs and reference compounds were as follows, with higher concentrations for methods involving OH radical scavengers: 7.87 × 1013 for 2-methyl-3-buten-2-ol, (5.47–10.94) × 1013 for 3-methyl-2-buten-1-ol, 7.86 × 1013 for 3-methyl-3-buten-1-ol, 7.69 × 1013 for 2-methyl-3-buten-1-ol, (7.90–12.73) × 1013 for 3-methyl-3-buten-2-ol, 3.98 × 1013 for 2-methyl-2-butene, (3.98–7.96) × 1013 for cyclohexene, (6.49–16.21) × 1013 for propene, (6.49–8.11) × 1013 for ethene, (6.49–25.94) × 1013 for E-2-butene, 6.49 × 1013 for 3-methyl-1-butene, and 6.49 × 1013 for 1-butene. The concentration of (4.78–7.96) × 1013 for 1,3,5-trimethylbenzene was used when this compound was employed as a tracer for the OH radicals, while the concentrations of CO and 1,3,5-trimethylbenzene acting as scavengers for OH radicals exhibited a range of values between (2.69–5.38) × 1017 and (8.55–28.49) × 1014, respectively.
2.1. Relative Rate Kinetic Setup
The processes occurring within the reactor during the relative kinetic measurements are described in reactions 6–8.
| 6 |
| 7 |
| 8 |
Provided that the methyl-butenols and the reference compounds react exclusively with ozone in the presence of an OH scavenger, eq I can be used to describe the relationship between MBO decay and reference decay.
| I |
where [MBO]t0, [MBO]t, [Ref]t0, and [Ref]t are the concentrations of the MBO and the reference compound at t0 and t, respectively, and k1, k2, and k3 are the rate coefficients corresponding to reactions 6–8. Consequently, plots of ln([MBO]t0/[MBO]t)– k3 × (t – t0) against ln([reference]t0/[reference]t) should yield straight lines with zero intercept and the slope of k1/k2.
The relative kinetic method was employed to determine the MBO + O3 reaction rate coefficients. The reference compounds used in this study were ethene, propene, 1-butene, 3-methyl-1-butene, E-2-butene, 2-methyl-2-butene, and cyclohexene. The rate coefficient values of the reference compounds with O3 as used in the present study (in cm3 molecule–1 s–1) were kethene = (1.55 ± 0.23) × 10–18,34kpropene = (1.01 ± 0.15) × 10–17,35k1-butene = (9.65 ± 1.45) × 10–18,36k3-methyl-1-butene = (7.30 ± 1.10) × 10–18,37 kE-2-butene = (2.00 ± 0.30) × 10–16,34k2-methyl-2-butene = (4.03 ± 0.60) × 10–16,38 and kcyclohexene = (8.10 ± 1.22) × 10–17.39
The concentrations of CO and 1,3,5-trimethylbenzene were adjusted to achieve a scavenging efficiency of at least 90% (y = 0.9) of the OH radicals generated during ozonolysis. The initial scavenger concentration ([Scav]t0) was calculated in accordance with eq II, using the gas-phase reaction rate coefficients (k(A+OH)) toward OH radicals and the initial concentration ([A]t0) of the employed analytes (alkenols and olefins) as well as the gas-phase rate coefficients (k(Scav+OH)) of the corresponding OH radical reaction with the selected scavenger compound. The scavenger rate coefficient values used (in cm3 molecule–1 s–1) for the OH radical reaction were (2.40 ± 0.12) × 10–13 for carbon monoxide35 and (5.86 ± 0.88) × 10–11 for 1,3,5-trimethylbenzene.34
| II |
Another method for assessing the reaction rate coefficients of volatile organic compounds with ozone is to use a tracer compound to evaluate the concentration of the OH radicals formed during ozonolysis and to correct for the decay of MBOs and the associated references.40,41 A suitable tracer compound should react with OH radicals but not with ozone molecules. In the present study, 1,3,5-trimethylbenzene was used as the tracer compound. In addition to reactions 6–8, 9–12 also occur in the gas mixture reactions when a tracer is used.
| 9 |
| 10 |
| 11 |
| 12 |
Monitoring the sink rate of the tracer, with respect to its gas-phase reaction rate coefficient, enables determination of the OH radical concentration ([OH]) as expressed in eq III.
| III |
where [T]t0 and [T]t represent the tracer concentration at time t0 and t, respectively, k4 is the wall loss rate of the tracer on reactor walls, and k′T represents the OH radical gas-phase reaction rate coefficient with the tracer compound at the working temperature and pressure conditions. Equation IV accounts for the secondary chemical reactions initiated by the OH radicals in the gas-phase mixture and is derived from eq I.
| IV |
where k′1 and k′2 are the gas-phase OH-initiated reaction rate coefficients of MBOs and of reference compounds, respectively. Therefore, the slope of (ln([MBO]t0/[MBO]t) – (k3 + k′1 × [OH]) × (t – t0)) against (ln([Ref]t0/[Ref]t) – k′2 × [OH] × (t – t0)) represents the ratio between the reaction rate coefficients of ozone with MBO and the reference compound (k1/k2).The OH reaction rate coefficients for the investigated MBOs were evaluated in a previous study19 and are as follows (in 10–11 cm3 molecules–1 s–1): kMBO232 = (6.32 ± 0.49), kMBO321 = (14.55 ± 0.93), kMBO331 = (10.04 ± 0.78), kMBO231 = (5.31 ± 0.37), and kMBO332 = (11.71 ± 1.29). For the reference compounds involved in the present study, the following OH radical-initiated gas-phase reaction rate coefficients (in units of 10–11 cm3 molecules–1 s–1) were used: kethene = (0.85 ± 0.13),34kpropene = (2.63 ± 0.39),36k3-methyl-1-butene = (3.18 ± 0.62),36kE-2-butene = (6.31 ± 0.95),34 and kcyclohexene = (6.77 ± 1.67).38
The chemicals were used as received without any purification: 2-methyl-3-buten-2-ol (Aldrich, 98%), 3-methyl-2-buten-1-ol (Aldrich, 99%), 3-methyl-3-buten-1-ol (Aldrich, 97%), 2-methyl-3-buten-1-ol (Aldrich, 98%), 3-methyl-3-buten-2-ol (Aldrich), ethene (Sigma-Aldrich, 99,95%), propene (Aldrich, 99.5%), E-2-butene (Aldrich, 99%), 1-butene (Sigma-Aldrich, >99.38%), 3-methyl-1-butene (Sigma-Aldrich, 95%), 2-methyl-2-butene (Sigma-Aldrich, >95%), cyclohexene (Sigma-Aldrich, 99%), 1,3,5-trimethylbenzene (Aldrich, 98%), CO (99.997%, Linde), synthetic air (Messer-Griesheim, 99.999%), and oxygen (Messer-Griesheim, 99.999%).
2.2. Structure–Activity Relationship (SAR) Methodologies
The estimation of the gas-phase reaction rate coefficients of ozone with unsaturated hydrocarbons is of paramount importance for the prediction of the atmospheric fate and impact of such compounds in the absence of experimental data. In this study, five SAR methodologies were used to estimate the reaction rate coefficients of ozone with the investigated methyl-butenols.
The Atkinson and Carter42 method involves the ethene structure (R2C = CR2) as the base reactivity coefficient in the calculations of the ozone reaction rate coefficient, with various substitution factors to accommodate the effects of functional groups upon the double bond.
The methodology proposed by Calvert et al.43 is based on the structural characteristics of polyalkenes and focuses exclusively on the individual interactions between ozone molecules and various substituted double bonds, while neglecting the effect of substituents.
The method of Pfrang et al.44 uses algorithms which are based on the alkene structure and the correlation between the experimental kinetic data and the energy of the highest occupied molecular orbital (HOMO) to estimate the reaction rate coefficients of alkenols and unsaturated ethers.
The method of McGillen et al.45 for estimating the gas-phase ozone-initiated oxidation of heteroatomic unsaturated volatile organic compounds (HUVOCs) accounts for the combination of inductive and steric effects occurring in various classes of atmospheric oxygenates. The steric effect factor is based solely on the molecular geometric disposition and a specific set of four rules derived from the previous SAR methodology46 on alkenes. The total inductive effect is calculated using optimized inductive constants corresponding to different substituents and accounts for their distance and branching degree between the substitution and the olefinic bond.
The Jenkin et al.47 method, with access to a larger experimental kinetic dataset, proposes a SAR algorithm that accounts for the degree of substitution of the double bonds, the type of substituents, and Arrhenius parameters. In contrast to the methodology proposed by McGillen et al.,45 the latest approach47 accounts solely for the effects of substituents at the α-carbon atom, while the steric effect is discretely included in the Fα(298)(X) parameters.
3. Results and Discussion
3.1. Rate Coefficients of O3 Reactions with the Unsaturated Alcohols
The second-order rate coefficients for reactions 1–5 were obtained from the relative loss of methyl-butenols versus the loss of the reference compounds in the presence of ozone. The kinetic plots for all of the investigated unsaturated alcohols were generated using both the scavenger and the tracer methods, as shown in Figures 1–5. The wall loss of the organic reactants of interest was taken into account for each particular case. The experimental data related to the wall loss estimation (measuring with FT-IR the behavior of the reaction mixture in the dark for 10 min) showed that MBO321, MBO331, and 1,3,5-trimethylbenzene are susceptible to this kind of sink processes. The measured rates (k3 or k4 ± 2σ) are (1.21 ± 0.07) × 10–4 s–1 for MBO321, (3.75 ± 0.28) × 10–5 s–1 for MBO331, and (3.15 ± 0.19) × 10–5 s–1 for trimethylbenzene.
Figure 1.
Relative kinetic plots of the ozone reaction with 3-methyl-2-buten-1-ol versus (●/○) cyclohexene, (◆/◇) E-2-butene, and (▼) 2-methyl-2-butene (filled symbols and continuous regression lines are data from the OH scavenger radical method; empty symbols and dashed regression lines are data from the OH radical tracer method). The terms k′1 × [OH] and k′2 × [OH] × (t – t0) were neglected when the scavenger has been used.
Figure 5.
Relative kinetic plots of the ozone reaction with 3-methyl-3-buten-2-ol versus (□) propene and (○) cyclohexene.
Figure 2.
Relative kinetic plots of the ozone reaction with 2-methyl-3-buten-2-ol versus (□) propene and (Δ) 3-methyl-1-butene.
Figure 3.
Relative kinetic plots of the ozone reaction with 3-methyl-3-buten-1-ol versus (■/□) propene, (★) 1-butene, and (▲/Δ) 3-methyl-1-butene (filled symbols and continuous regression lines are data from the OH scavenger radical method; empty symbols and dashed regression lines are data from the OH radical tracer method). The terms k′1 × [OH] and k′2 × [OH] × (t – t0) were neglected when the scavenger has been used. Note: to provide a more accurate representation of the data, the values for 1-butene and 3-methyl-1-butene have been shifted up on the y-axis by 0.1 and 0.2 units, respectively.
Figure 4.
Relative kinetic plots of the ozone reaction with 2-methyl-3-buten-1-ol versus (■/□) propene. and (⬟/⬠) ethene (filled symbols and continuous regression lines are data from the OH scavenger radical method; empty symbols and dashed regression lines are data from the OH radical tracer method). The terms k′1 × [OH] and k′2 × [OH] × (t – t0) were neglected when the scavenger has been used.
All plots in Figures 1–5 exhibit excellent correlation coefficients with near-zero intercepts. It should be noted that no significant processes driven by the OH radicals occur when a scavenger is used. Corrections due to secondary OH reactions of analytes are effective when using 1,3,5-trimethylbenzene as an OH tracer.
Tables 1 and 2 present the relative rate ratios (k1/k2 ± 2σ) and the rate coefficients (k1,i) determined for the investigated reactions while using a scavenger and tracer, respectively. Uncertainties for k1,i were calculated using the propagation of 2σ from the linear regression analysis of the experimental data with an additional 15% uncertainty from the reference compound reaction rate coefficients. Additionally, the k1(average) value and its associated uncertainty were calculated using a weighted average of the k1,i values, as previously described.19 The proposed rate coefficient for the investigated reactions is a weighted average of the k1,i values, as no significant differences were observed while using different reference compounds or different methods to mitigate OH radical interference.
Table 1. Rate Constants for the Reaction of Ozone with Methyl-butenols at Room Temperature, Obtained under Conditions where OH Radicals Were Scavenged.

Table 2. Rate Constants for the Reaction of Ozone with Methyl-Butenols at Room Temperature, Employing Tracers for OH Radicals and Overall Average Values for Rate Constants with Compounds Studied by Both Methods.

The OH radical concentration, calculated from the sink of 1,3,5-trimethylbenzene during the ozonolysis of the reaction mixture, ranged between (0.51 and 4.02) × 106 radicals cm–3. The ozone concentration within the reactor exhibited a range of values between (2.32 and 78.89) × 1012 molecules cm–3 during the course of the experiments. Given that the OH radical generation is directly related with the ozone concentration, in the present study, the corrections due to secondary chemistry initiated by the OH radicals when using the tracer method reach an upper limit of 29.8% for MBO232, 33.2% for MBO321, 47.1% for MBO331, 29.7% for MBO231, and 30.0% for MBO332 relative to the total decay of MBOs.
The present study is using two different methods found in the literature38,40,41 to determine the gas-phase reaction rate coefficients for the ozone-initiated oxidation of MBOs. The method involving an OH radical tracer is frequently used to estimate the concentration of the OH radicals available to initiate chemical processes into the gas phase.48 In the absence of a more accurate but expensive and laborious techniques like laser-induced fluorescence (LIF), fluorescence assay by gas expansion (FAGE), and chemical ionization mass spectrometry (CIMS), this method could provide consistent results.49
Both methods provide reliable data and can be used for rate coefficient determinations in ozonolysis reactions. The scavenger method is commonly used in kinetic studies, as it limits interferences from OH radical reactions. The presence of OH radicals and their reactions with reactants generate products that complicate the evaluation of the kinetic results. However, the necessity to eliminate over 90% of the OH radicals requires very high scavenger concentrations, which can hinder monitoring techniques.
On the other hand, the tracer technique, while maintaining low reaction mixture concentrations, does not eliminate OH radicals produced in ozonolysis. Instead, it allows quantification of OH radical concentration further used for the corrections of the additional reactant loss in secondary reactions involving OH chemistry. The FTIR technique used in this study allowed comparison of both methods within the same experimental setup.
The relative kinetic plots of reference versus reference are presented in Figures S1 and S2 in the Supporting information (SI). These control tests reveal excellent correlations between the reference reaction rate coefficients considered within the present study. Differences lower than 10% were obtained from the statistical analysis performed on this dataset (Table S1 in the SI).
A comparison of the obtained reaction rate coefficients with the available literature data is presented in Table 3. The rate coefficient for MBO232 of (9.55 ± 1.04) × 10–18 cm3 molecule–1 s–1 is within 11% of the relative rate coefficient reported by Fantechi et al.28 at room temperature. The aforementioned rate coefficient is consistent with the rate constants reported by Grosjean and Grosjean27 and Klawatsch-Carrasco et al.29 who employed absolute rate methods in the performed studies. Grosjean and Grosjean30 reported a value of (439 ± 37) × 10–18 cm3 molecule–1 s–1 for MBO321, which is 40% higher than the value of (311 ± 20) × 10–18 cm3 molecule–1 s–1 obtained in this study. The value reported by Grosjean and Grosjean30 was obtained in a single experiment without any reproducibility checks. It was determined by monitoring the ozone decay using ultraviolet photometry, for which the authors infer that it could interfere with other unsaturated oxygenates. Additionally, the humidity conditions (RH = 55%) specific for Grosjean and Grosjean30 work could accelerate the decomposition of the MBO–O3 complex.50 For MBO331 and MBO231, the obtained values are (7.29 ± 0.46) and (4.25 ± 0.29) × 10–18 cm3 molecule–1, respectively. The values obtained in this study are in agreement with those reported by Gai et al.31 of (6.80 ± 1.29) and (3.74 ± 0.62) × 10–18 cm3 molecule–1 s–1, respectively. To date, no information regarding the reaction rate coefficient of MBO332 with ozone is available in the literature.
Table 3. Rate Coefficients Obtained in the Present Work for the Reaction between Methyl-Butenols and Ozone, with a Comparison to the Existing Literature Dataa.
AR: absolute rate method; RR: relative rate method.
3.2. Comparison between Experimental Data and SAR Estimates
The ozone gas-phase reaction rate estimates were obtained using five different methodologies.42−45,47 The results are presented in Table 4 along with the experimental kinetic values obtained within the present study. The calculation procedure is presented in Tables S2–S5 in the SM. The least-squared fit statistical analysis of the estimates with respect to the experimental reaction rate coefficients suggests that the method of McGillen et al.45 performs better than the others (∑((kMBO – kSAR)/kMBO)2 = 0.55). The second-best fit was observed for the method of Calvert et al.43 (∑((kMBO – kSAR)/kMBO)2 = 0.60), followed by the Jenkin et al.47 methodology (∑((kMBO– kSAR)/kMBO)2 = 0.64), Atkinson and Carter42 (∑((kMBO – kSAR)/kMBO)2 = 0.78), and Pfrang et al.44 (∑((kMBO – kSAR)/kMBO)2 = 0.95).
Table 4. Comparison of the Rate Coefficients for the Reaction of Ozone with the Methyl-butenols Investigated in This Work Employing Different SAR Methods.
Although there are differences between SAR approaches, these can be actually dependent on the way the SAR automated construction mechanism was developed. The SAR methodology of Calvert et al.43 accounts only for the substitution degree of the olefinic bond(s). Furthermore, the Atkinson and Carter42 method employs specific scaling factors to accommodate the substitution degree and functional groups, although it lacks the factor corresponding to the OH alcohol group.
Pfrang et al.44 proposed a methodology that failed to consider steric hindrance by correlating the energy of the highest occupied molecular orbital with the rate coefficients. This methodology assumes that higher chain substituents and further positioning of the OH functional group from the olefinic bond would increase the reaction rate coefficients.
The method of Jenkin et al.47 estimates ozone reaction rate coefficients, emphasizing the substitution degree of (poly)alkenes, geometrical isomerization, Arrhenius parameters, and the nature of β-substituents, while also neglecting the γ-substituent effects and steric hindrances. McGillen et al.45 accounts for the steric hindrance of the substituent and differences occurring between the positioning and branching degree when calculating the total inductive effect of various functionalities. The model of McGillen et al.45 proved to be the most suitable methodology for estimating aliphatic functionalized unsaturated volatile reaction rate coefficients toward ozone molecules under ambient conditions.
3.3. Substituent Effects on Reactivity
The experimental kinetic data presented in Table 2 clearly demonstrate that the reactivity trend in the gas phase toward ozone molecules is consistent with the following order: kMBO231 < kMBO331 < kMBO232 < kMBO332 < kMBO321. It is widely acknowledged that the ozone reaction rate coefficients increase with the degree of substitution of the double bond, as exemplified by the case of MBO321, which possesses a trisubstituted unit. The reaction of ozone with unsaturated alcohols in the gas phase involves the electrophilic addition of ozone to the unsaturated carbon–carbon bond, followed by unimolecular decomposition of the ozonide into carbonyls and Criegee intermediates.38 A comparative reactivity analysis of the MBO331 versus MBO332 and of the MBO231 versus MBO232 reveals a notable increase in their reactivity toward ozone for the last MBOs. The alcohol functional group can exert either an activating effect, by stabilizing the ozonide when situated at the carbon atom adjacent to the double bond,27 or a steric hindering effect, preventing ozone from reaching the double bond when situated in the γ or δ position relative to the olefinic bond. The stabilizing effect of the OH functional group should diminish with increasing temperature, as evidenced by a comparison of the Arrhenius parameters of alkenols and their homologue alkenes. An example is provided by a comparison of the kO3 variation of 1-penten-3-ol, cis-2-penten-1-ol, and trans-3-hexen-1-ol51 with that of 1-pentene, cis-2-pentene, and trans-3-hexene,52 respectively. A lower increase in reactivity toward ozone molecules is observed for alkenols, with increasing temperature, especially monosubstituted alkenols. This suggests that the OH alcohol group may physically interact with the trioxolane cycle of the ozonide via a H donor/acceptor interaction.
For comparison purposes, the ozone reaction rate coefficients of the investigated unsaturated alcohols and their alkene and carbonyl structural homologues are provided in Table 5. The presence of the OH functional group (i) in the β position to the olefinic bond increases the reactivity of mono- and disubstituted alkenols, as evidenced by the reactivity of MBO232 and MBO332 in comparison with the reactivity of 3-methyl-1-butene and 2-methyl-1-butene, respectively, and (ii) decreases when the OH functional group is situated in the γ position, as observed for MBO231 and MBO331. The reactivity of trisubstituted alkenols appears to be lower than that of homologue alkenes, as evidenced by the case of MBO321 and 2-methyl-2-butene. A difference of 2 orders of magnitude can be observed when comparing the reactivity of MBO321 and its corresponding carbonyl derivative, 3-methyl-2-butenal, toward ozone. Additionally, a notable reduction in reactivity is observed when comparing MBO332 and 3-methyl-3-buten-2-one. The observed decrease in reactivity toward ozone can be attributed to the steric hindrance of the alcohol function in MBO321, MBO331, and MBO231, as evidenced by the kinetic data available for carbonyl derivatives and also for higher chain alkenols.21,22,53
Table 5. Rate Coefficients (in Units of cm3 molecule–1 s–1) for the Gas-Phase Reaction of Ozone with Methyl-butenols and Their Alkene and Carbonyl Structural Homologues.
| methyl-butenol | kO3 × 1018a | alkenes | kO3 × 1018 | kMBO/kalkene | carbonyls | kO3 × 1018 | kMBO/kCarbonyl |
|---|---|---|---|---|---|---|---|
| 3-methyl-2-buten-1-ol | 311 | 2-methyl-2-butene | 40338 | 0.77 | 3-methyl-2-butenal | 1.8254 | 171 |
| 3-methyl-3-buten-1-ol | 7.29 | 2-methyl-1-butene | 14.3052 | 0.51 | |||
| 3-methyl-3-buten-2-ol | 62.9 | 4.40 | 3-methyl-3-buten-2-one | 11.855 | 5.33 | ||
| 2-methyl-3-buten-1-ol | 4.25 | 3-methyl-1-butene | 7.3037 | 0.58 | |||
| 2-methyl-3-buten-2-ol | 9.55 | 1.30 |
This work.
3.4. Atmospheric implications
The average lifetimes (τ = 1/(∑(kox × [ox]))) of the investigated MBOs due to reactions with atmospherically relevant oxidants were calculated and are provided in Table 6. The calculated lifetimes for the MBOs due to the tropospheric ozone reaction were as follows: 1.3 h for MBO321, 6.3 h for MBO332, 41.6 h for MBO232, 54.4 h for MBO331, and 93.4 h for MBO231. MBO321 and MBO332 have atmospheric lifetimes with respect to the reaction with OH and ozone, which are similar; the other MBOs are lost predominantly via OH radicals during the daytime.19
Table 6. Updated Tropospheric Lifetimes, τ (in Hours), for the Methyl-butenol Series with OH, O3, NO3, and Cl Atmospheric Oxidant Speciesa,b.
|
rate
coefficient(cm3 × molecule–1 × s–1) |
tropospheric lifetime (h) |
||||||||
|---|---|---|---|---|---|---|---|---|---|
| methyl-butenol | kOH × 1011 | kO3 × 1018a | kNO3 × 1013 | kCl × 1010 | τOH | τO3 | τNO3 | τCl | τtotal |
| 2-methyl-3-buten-2-ol | 6.3219 | 9.55 | 0.1156 | 2.6457 | 3.89 | 41.55 | 50.51 | 105.22 | 3.22 |
| 3-methyl-2-buten-1-ol | 14.5519 | 311 | 10.0056 | 4.0258 | 1.69 | 1.27 | 0.56 | 69.1 | 0.31 |
| 3-methyl-3-buten-1-ol | 10.0419 | 7.29 | 2.7056 | 4.1331 | 2.45 | 54.4 | 2.06 | 67.3 | 1.08 |
| 2-methyl-3-buten-1-ol | 5.3119 | 4.25 | (−) | 3.5131 | 4.63 | 93.4 | (−) | 79.1 | <4.18 |
| 3-methyl-3-buten-2-ol | 11.7119 | 62.9 | (−) | (−) | 2.10 | 6.31 | (−) | (−) | <1.58 |
The photochemical ozone creation potential (POCPE) index has been calculated for the investigated alkenols using the method proposed by Jenkin et al.32 based on the gas-phase OH-reaction rate coefficients19 and the O3-reaction rate coefficients determined within the present study. This method quantitatively estimated the abilities of VOCs to produce tropospheric ozone and accounts for both the multiday regional-scale ozone formation typical for northwestern European conditions and the single-day ozone formation typical of urban conditions in the United States. The obtained POCPE index for northwestern European conditions was estimated to be as follows: 93 for MBO321, 114 for MBO231, 119 for MBO232, 127 for MBO331, and 140 for MBO332. The POCPE index values should be interpreted in comparison to the POCPE value, equal to unity for ethene. The POCPE index for USA urban conditions was found to be 155 for MBO231, 190 for MBO232, 238 for MBO331, 783 for MBO332, and 5346 for MBO321. The differences in the POPCE index of MBO321 and MNO332 under high-NOx conditions can be attributed to their high reactivity toward OH radicals in comparison with other isomers. This reactivity subsequently leads to the formation of RO2-peroxy radicals, which shifts the atmospheric balance toward NO2 formation. In low-NOx conditions, the RO2-type radicals will deactivate each other without the formation of O3 precursors, thus lowering the POCPE index value.
4. Conclusions
The gas-phase reaction rate coefficients of five methyl-butenols with ozone molecules were experimentally determined by a relative rate approach at a temperature of (298 ± 2) K and total air pressure of (1000 ± 10) mbar using both scavenger and tracer methods. This work represents the first reported experimental determination of the rate coefficient for the reaction between 3-methyl-3-buten-2-ol and ozone. The experimental data obtained by the two different methods, using at least two reference compounds, demonstrated a strong correlation and credibility for the reaction rate coefficients for the MBO series with the ozone molecule. The associated interferences have been properly corrected; thus, consistent reaction rate coefficients were obtained employing those methods for MBO321, MBO331, and MBO231.
The experimental data indicate an increase in reactivity toward ozone when the OH functional group is situated in the β position to the double bond. This observation is consistent when comparing homologue alkenes. The establishment of an interaction between the alkenol–O3 complex and the alcohol functional might be responsible for this phenomenon. Specific characteristics emerge when γ-functional groups can sterically hinder the double bond. This results in a decrease of the reaction rate coefficients for MBO331 and MBO231 compared with the alkene series. Additionally, the degree of substitution plays a significant role in the reactivity trend.
The SAR methodology presented by McGillen et al.45 demonstrates the most effective correlation between estimates and the experimental kinetic data presented in the current study.
Atmospheric lifetimes were estimated for the investigated compounds. It was found that the atmospheric reaction triggered by ozone molecules was as important as the reactions initiated by the OH radicals for MBO321 and MBO332. For all other MBOs, the main atmospheric sink is represented by the reaction with OH radicals.
The dependence of the ozone reaction rate of alkenols on temperature and relative humidity may further elucidate crucial aspects of the atmospheric behavior of such compounds, from which SAR and global atmospheric distribution models could benefit from. Moreover, mechanistic studies conducted under relevant atmospheric conditions are essential to assess the products, aerosols, and photochemical formation during the sinking of aliphatic unsaturated oxygenates.
Acknowledgments
The authors acknowledge the financial support from PN-III-P4-PCE2021-0673 (acronym ATMO-SOS).The authors acknowledge the Operational Program Competitiveness 2014-2020, Axis 1, under POC/448/1/1 research infrastructure projects for public R&D institutions/sections F 2018, through the Research Center with Integrated Techniques for Atmospheric Aerosol Investigation in Romania (RECENT AIR) project, under grant agreement MySMIS no. 127324, and the European Commission under the Horizon 2020 Research and Innovation Framework Programme, H2020-INFRAIA-2020, Grant Agreement number 101008004 (ATMO-ACCESS).
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpca.4c03653.
Table S1: Experimentally determined rate constants for the reaction of ozone with the employed reference compounds at (298 ± 2) K and (1000 ± 10) mbar in air buffer gas. Table S2: Estimates for the gas-phase reaction of ozone with selected methyl-butenols at 298 K calculated according to the Atkinson and Carter and Calvert et al. SAR methodology for alkenes. Table S3: Estimates for the gas-phase reaction of ozone with selected methyl-butenols at 298 K calculated according to the Pfrang et al. SAR methodology for alkenols. Table S4: Estimates for the gas-phase reaction of ozone with selected methyl-butenols at 298 K calculated according to the McGillen et al. SAR methodology for HUVOCs. Table S5: Estimates for the gas-phase reaction of ozone with selected methyl-butenols at 298 K calculated according to the Jenkin et al. SAR methodology for unsaturated organic compounds. Figure S1: Control kinetic plots of 2-methyl-2-butene and cyclohexene versus E-2-butene and of 1-butene, 3-methyl-1-butene, and ethene versus propene during the ozonolysis of MBOs in the presence of of an OH scavenger. Figure S2: Control kinetic plots of cyclohexene versus E-2-butene and of cyclohexene, 3-methyl-1-butene, and ethene versus propene during the ozonolysis of MBOs in the presence of an OH tracer (PDF)
Author Contributions
∥ A.-M.R. and C.R. authors with equal contribution.
The authors declare no competing financial interest.
Supplementary Material
References
- Koppmann R.Volatile Organic Compounds in the Atmosphere; Koppmann R., Ed.; Wiley: Oxford: UK, 2007. [Google Scholar]
- Blande J. D.; Holopainen J. K.; Niinemets Ü. Plant Volatiles in Polluted Atmospheres: Stress Responses and Signal Degradation. Plant. Cell Environ. 2014, 37 (8), 1892–1904. 10.1111/pce.12352. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calvert J.; Mellouki A.; Orlando J.; Pilling M.; Wallington T.. Mechanisms of Atmospheric Oxidation of the Oxygenates; Oxford University Press, 2011. [Google Scholar]
- Hellpointner E.; Gäb S. Detection of Methyl, Hydroxymethyl and Hydroxyethyl Hydroperoxides in Air and Precipitation. Nature 1989, 337 (6208), 631–634. 10.1038/337631a0. [DOI] [Google Scholar]
- Hallquist M.; Wenger J. C.; Baltensperger U.; Rudich Y.; Simpson D.; Claeys M.; Dommen J.; Donahue N. M.; George C.; Goldstein A. H.; et al. The Formation, Properties and Impact of Secondary Organic Aerosol: Current and Emerging Issues. Atmos. Chem. Phys. 2009, 9 (14), 5155–5236. 10.5194/acp-9-5155-2009. [DOI] [Google Scholar]
- Kanakidou M.; Seinfeld J. H.; Pandis S. N.; Barnes I.; Dentener F. J.; Facchini M. C.; Van Dingenen R.; Ervens B.; Nenes A.; Nielsen C. J.; et al. Organic Aerosol and Global Climate Modelling: A Review. Atmos. Chem. Phys. 2005, 5 (4), 1053–1123. 10.5194/acp-5-1053-2005. [DOI] [Google Scholar]
- Lelieveld J.; Evans J. S.; Fnais M.; Giannadaki D.; Pozzer A. The Contribution of Outdoor Air Pollution Sources to Premature Mortality on a Global Scale. Nature 2015, 525 (7569), 367–371. 10.1038/nature15371. [DOI] [PubMed] [Google Scholar]
- 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. Phys. Chem. 2010, 224 (7–8), 1059–1080. 10.1524/zpch.2010.6141. [DOI] [Google Scholar]
- Gray D. W. Transient Increases in Methylbutenol Emission Following Partial Defoliation of Pinus Ponderosa. Atmos. Environ. 2008, 42 (19), 4797–4802. 10.1016/j.atmosenv.2008.01.048. [DOI] [Google Scholar]
- Lerdau M.; Gray D. Ecology and Evolution of Light-dependent and Light-independent Phytogenic Volatile Organic Carbon. New Phytol. 2003, 157 (2), 199–211. 10.1046/j.1469-8137.2003.00673.x. [DOI] [PubMed] [Google Scholar]
- Lehnert A.-S.; Perreca E.; Gershenzon J.; Pohnert G.; Trumbore S. E. Simultaneous Real-Time Measurement of Isoprene and 2-Methyl-3-Buten-2-Ol Emissions From Trees Using SIFT-MS. Front. Plant Sci. 2020, 11, 578204. 10.3389/fpls.2020.578204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schade G. W.; Goldstein A. H. Fluxes of Oxygenated Volatile Organic Compounds from a Ponderosa Pine Plantation. J. Geophys. Res.: Atmos. 2001, 106 (D3), 3111–3123. 10.1029/2000JD900592. [DOI] [Google Scholar]
- 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 (8), 861–874. 10.1016/1352-2310(95)00026-U. [DOI] [Google Scholar]
- Zhang Q.-H.; Schlyter F.; Birgersson G. 2-Methyl-3-Buten-2-Ol: A Pheromone Component of Conifer Bark Beetles Found in the Bark of Nonhost Deciduous Trees. Psyche: J. Entomol. 2012, 2012 (1), 414508. 10.1155/2012/414508. [DOI] [Google Scholar]
- Zhao T.; Ganji S.; Schiebe C.; Bohman B.; Weinstein P.; Krokene P.; Borg-Karlson A.-K.; Unelius C. R. Convergent Evolution of Semiochemicals across Kingdoms: Bark Beetles and Their Fungal Symbionts. ISME J. 2019, 13 (6), 1535–1545. 10.1038/s41396-019-0370-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bernard F.; Daële V.; Mellouki A.; Sidebottom H. Studies of the Gas Phase Reactions of Linalool, 6-Methyl-5-Hepten-2-Ol and 3-Methyl-1-Penten-3-Ol with O3 and OH Radicals. J. Phys. Chem. A 2012, 116 (24), 6113–6126. 10.1021/jp211355d. [DOI] [PubMed] [Google Scholar]
- 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 (7), 3347–3358. 10.5194/acp-11-3347-2011. [DOI] [Google Scholar]
- Gibilisco R. G.; Blanco M. 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 (13), 7717–7725. 10.1021/es506125c. [DOI] [PubMed] [Google Scholar]
- Rusu Vasilache A.-M.; Roman C.; Bejan I. G.; Arsene C.; Olariu R. I. Gas-Phase Kinetic Investigation of the OH-Initiated Oxidation of a Series of Methyl-Butenols under Simulated Atmospheric Conditions. J. Phys. Chem. A 2024, 128 (24), 4838–4849. 10.1021/acs.jpca.4c02287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allani A.; Romanias M. N. Reassessment of the Temperature Dependent Oxidation of 2-methyl-3-butene-2-ol (MBO) by Cl Atoms: A Kinetic and Product Study. Int. J. Chem. Kinet. 2022, 54 (7), 424–434. 10.1002/kin.21571. [DOI] [Google Scholar]
- Grira A.; Amarandei C.; Roman C.; Bejaoui O.; Aloui N.; El Dib G.; Arsene C.; Bejan I. G.; Olariu R. I.; Canosa A.; Tomas A. Gas-Phase Ozone Reaction Kinetics of C 5 – C 8 Unsaturated Alcohols of Biogenic Interest. J. Phys. Chem. A 2022, 126 (27), 4413–4423. 10.1021/acs.jpca.2c02805. [DOI] [PubMed] [Google Scholar]
- 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. 10.1016/j.cplett.2014.11.003. [DOI] [Google Scholar]
- Xu Y.; Tong S.; Li W.; Chen M.; Hu L.; Zhang H.; Wang S.; Ge M. Nighttime Reactions of a Series of Unsaturated Alcohols with NO3•: Kinetics, Products and Mechanisms Study. J. Environ. Sci. 2025, 151, 331–346. 10.1016/j.jes.2024.03.043. [DOI] [Google Scholar]
- Molteni U.; Simon M.; Heinritzi M.; Hoyle C. R.; Bernhammer A.-K.; Bianchi F.; Breitenlechner M.; Brilke S.; Dias A.; Duplissy J.; et al. Formation of Highly Oxygenated Organic Molecules from α-Pinene Ozonolysis: Chemical Characteristics, Mechanism, and Kinetic Model Development. ACS Earth Space Chem. 2019, 3 (5), 873–883. 10.1021/acsearthspacechem.9b00035. [DOI] [Google Scholar]
- Keene W. C.; Galloway J. N.; Holden J. D. Measurement of Weak Organic Acidity in Precipitation from Remote Areas of the World. J. Geophys. Res.: Oceans 1983, 88 (C9), 5122–5130. 10.1029/JC088iC09p05122. [DOI] [Google Scholar]
- Geyer A.; Bächmann K.; Hofzumahaus A.; Holland F.; Konrad S.; Klüpfel T.; Pätz H.; Perner D.; Mihelcic D.; Schäfer H.. et al. Nighttime Formation of Peroxy and Hydroxyl Radicals during the BERLIOZ Campaign: Observations and Modeling Studies J. Geophys. Res.: Atmos. 2003; Vol. 108 (D4), 10.1029/2001JD000656. [DOI]
- Grosjean D.; Grosjean E.; Williams E. L. Rate Constants for the Gas-Phase Reactions of Ozone with Unsaturated Alcohols, Esters, and Carbonyls. Int. J. Chem. Kinet. 1993, 25 (9), 783–794. 10.1002/kin.550250909. [DOI] [Google Scholar]
- Fantechi G.; Jensen N.; 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 (8), 589–594. . [DOI] [Google Scholar]
- 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 (3), 152–156. 10.1002/kin.10175. [DOI] [Google Scholar]
- Grosjean E.; Grosjean D. The Reaction of Unsaturated Aliphatic Oxygenates with Ozone. J. Atmos. Chem. 1999, 32 (2), 205–232. 10.1023/A:1006122000643. [DOI] [Google Scholar]
- Gai Y.; Ge M.; Wang W. Kinetics of the Gas-Phase Reactions of Some Unsaturated Alcohols with Cl Atoms and O3. Atmos. Environ. 2011, 45 (1), 53–59. 10.1016/j.atmosenv.2010.09.047. [DOI] [Google Scholar]
- Jenkin M. E.; Derwent R. G.; Wallington T. J. Photochemical Ozone Creation Potentials for Volatile Organic Compounds: Rationalization and Estimation. Atmos. Environ. 2017, 163, 128–137. 10.1016/j.atmosenv.2017.05.024. [DOI] [Google Scholar]
- Roman C.; Arsene C.; Bejan I. G.; Olariu R. I. Investigations into the Gas-Phase Photolysis and OH Radical Kinetics of Nitrocatechols: Implications of Intramolecular Interactions on Their Atmospheric Behaviour. Atmos. Chem. Phys. 2022, 22 (4), 2203–2219. 10.5194/acp-22-2203-2022. [DOI] [Google Scholar]
- McGillen M. R.; Carter W. P. L.; Mellouki A.; J Orlando J.; Picquet-Varrault B.; J Wallington T. Database for the Kinetics of the Gas-Phase Atmospheric Reactions of Organic Compounds. Earth Syst. Sci. Data 2020, 12 (2), 1203–1216. 10.5194/essd-12-1203-2020. [DOI] [Google Scholar]
- Burkholder J. B.; Sander S. P.; Abbatt J.; Barker J. R.; Cappa C.; Crounse J. D.; Dibble T. S.; Huie R. E.; Kolb C. E.; Kurylo M. J.. et al. Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies, Number 19JPL Public.Pasadena. 2019; Vol. http://jpldataeval.jpl.nasa.gov/.
- Atkinson R.; Arey J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103 (3), 4605–4638. 10.1021/cr0206420. [DOI] [PubMed] [Google Scholar]
- Shi Y.; Xu Y.; Jia L. Arrhenius Parameters for the Gas-Phase Reactions of O3 with Two Butenes and Two Methyl-Substituted Butenes over the Temperature Range of 295–351 K. Int. J. Chem. Kinet. 2011, 43 (5), 238–246. 10.1002/kin.20553. [DOI] [Google Scholar]
- Calvert J. G.; Orlando J. J.; Stockwell W. R.; Wallington T. J.. The Mechanisms of Reactions Influencing Atmospheric Ozone; Oxford University Press, 2015. [Google Scholar]
- Stewart D. J.; Almabrok S. H.; Lockhart J. P.; Mohamed O. M.; Nutt D. R.; Pfrang C.; Marston G. The Kinetics of the Gas-Phase Reactions of Selected Monoterpenes and Cyclo-Alkenes with Ozone and the NO3 Radical. Atmos. Environ. 2013, 70, 227–235. 10.1016/j.atmosenv.2013.01.036. [DOI] [Google Scholar]
- Paulson S. E.; Fenske J. D.; Sen A. D.; Callahan T. W. A Novel Small-Ratio Relative-Rate Technique for Measuring OH Formation Yields from the Reactions of O 3 with Alkenes in the Gas Phase, and Its Application to the Reactions of Ethene and Propene. J. Phys. Chem. A 1999, 103 (13), 2050–2059. 10.1021/jp984140v. [DOI] [Google Scholar]
- Rickard A. R.; Johnson D.; McGill C. D.; Marslon G. OH Yields in the Gas-Phase Reactions of Ozone with Alkenes. J. Phys. Chem. A 1999, 103 (38), 7656–7664. 10.1021/jp9916992. [DOI] [Google Scholar]
- Atkinson R.; Carter W. P. L. Kinetics and Mechanisms of the Gas-Phase Reactions of Ozone with Organic Compounds under Atmospheric Conditions. Chem. Rev. 1984, 84 (5), 437–470. 10.1021/cr00063a002. [DOI] [Google Scholar]
- Calvert J. G.; Atkinson R.; Kerr J. A.; Madronich S.; Moortgat G. K.; Wallington T. J.; Yarwood G.. The Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press, 2000. [Google Scholar]
- 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 (13), 3018–3034. 10.1016/j.atmosenv.2007.12.046. [DOI] [Google Scholar]
- 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 (7), 2842–2849. 10.1039/C0CP01732A. [DOI] [PubMed] [Google Scholar]
- McGillen M. R.; Carey T. J.; Archibald A. T.; Wenger J. C.; Shallcross D. E.; Percival C. J. Structure-Activity Relationship (SAR) for the Gas-Phase Ozonolysis of Aliphatic Alkenes and Dialkenes. Phys. Chem. Chem. Phys. 2008, 10 (13), 1757–1768. 10.1039/b715394e. [DOI] [PubMed] [Google Scholar]
- Jenkin M. E.; Valorso R.; Aumont B.; Newland M. J.; Rickard A. R. Estimation of Rate Coefficients for the Reactions of O3 with Unsaturated Organic Compounds for Use in Automated Mechanism Construction. Atmos. Chem. Phys. 2020, 20 (21), 12921–12937. 10.5194/acp-20-12921-2020. [DOI] [Google Scholar]
- Winiberg F. A. F.; Warman W. J.; Brumby C. A.; Boustead G.; Bejan I. G.; Speak T. H.; Heard D. E.; Stone D.; Seakins P. W. Comparison of Temperature-Dependent Calibration Methods of an Instrument to Measure OH and HO 2 Radicals Using Laser-Induced Fluorescence Spectroscopy. Atmos. Meas. Technol. 2023, 16 (19), 4375–4390. 10.5194/amt-16-4375-2023. [DOI] [Google Scholar]
- Heard D.Analytical Techniques for Atmospheric Measurement, 1st ed.; Heard D., Ed.; Blackwell Publishing: Oxford, 2006. [Google Scholar]
- Almatarneh M. H.; Elayan I. A.; Abu-Saleh A. A.; Altarawneh M.; Ariya P. A. The Gas-Phase Ozonolysis Reaction of Methylbutenol: A Mechanistic Study. Int. J. Quantum Chem. 2019, 119 (10), e25888. 10.1002/qua.25888. [DOI] [Google Scholar]
- 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 (2019), 117306 10.1016/j.atmosenv.2020.117306. [DOI] [Google Scholar]
- Avzianova E. V.; Ariya P. A. Temperature-dependent Kinetic Study for Ozonolysis of Selected Tropospheric Alkenes. Int. J. Chem. Kinet. 2002, 34 (12), 678–684. 10.1002/kin.10093. [DOI] [Google Scholar]
- 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 (87), 83573–83580. 10.1039/C6RA17107A. [DOI] [Google Scholar]
- Sato K.; Klotz B.; Taketsugu T.; Takayanagi T. Kinetic Measurements for the Reactions of Ozone with Crotonaldehyde and Its Methyl Derivatives and Calculations of Transition-State TheoryElectronic Supplementary Information (ESI) Available: The Stationary-Point Geometries Optimized at B3LYP/6-31G(d,p) F. Phys. Chem. Chem. Phys. 2004, 6 (15), 3969. 10.1039/b402496f. [DOI] [Google Scholar]
- Wang J.; Zhou L.; Wang W.; Ge M. Gas-Phase Reaction of Two Unsaturated Ketones with Atomic Cl and O 3: Kinetics and Products. Phys. Chem. Chem. Phys. 2015, 17 (18), 12000–12012. 10.1039/C4CP05461J. [DOI] [PubMed] [Google Scholar]
- Noda J.; Nyman G.; Langer S. Kinetics of the Gas-Phase Reaction of Some Unsaturated Alcohols with the Nitrate Radical. J. Phys. Chem. A 2002, 106 (6), 945–951. 10.1021/jp012329s. [DOI] [Google Scholar]
- 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 (12), 4224–4231. 10.1021/jp908104r. [DOI] [PubMed] [Google Scholar]
- Rodriguez D.; Rodriguez A.; Soto A.; Aranda A.; Diaz-De-Mera Y.; Notario A. Kinetics of the Reactions of Cl Atoms with 2-Buten-1-Ol, 2-Methyl-2-Propen-1-Ol, and 3-Methyl-2-Buten-1-Ol as a Function of Temperature. J. Atmos. Chem. 2008, 59 (3), 187–197. 10.1007/s10874-008-9101-6. [DOI] [Google Scholar]
- Lelieveld J.; Gromov S.; Pozzer A.; Taraborrelli D. Global Tropospheric Hydroxyl Distribution, Budget and Reactivity. Atmos. Chem. Phys. 2016, 16 (19), 12477–12493. 10.5194/acp-16-12477-2016. [DOI] [Google Scholar]
- Logan J. A. Tropospheric Ozone: Seasonal Behavior, Trends, and Anthropogenic Influence. J. Geophys. Res. 1985, 90 (D6), 10463–10482. 10.1029/JD090iD06p10463. [DOI] [Google Scholar]
- Shu Y.; Atkinson R. Atmospheric Lifetimes and Fates of a Series of Sesquiterpenes. J. Geophys. Res. 1995, 100 (D4), 7275–7281. 10.1029/95JD00368. [DOI] [Google Scholar]
- Wingenter O. W.; Kubo M. K.; Blake N. J.; Smith T. W.; Blake D. R.; Rowland F. S. Hydrocarbon and Halocarbon Measurements as Photochemical and Dynamical Indicators of Atmospheric Hydroxyl, Atomic Chlorine, and Vertical Mixing Obtained during Lagrangian Flights. J. Geophys. Res.: Atmos. 1996, 101 (D2), 4331–4340. 10.1029/95JD02457. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







