Abstract
Five biogenic unsaturated alcohols have been investigated under simulated atmospheric conditions regarding their gas-phase OH reactivity. The gas-phase rate coefficients of OH radicals with 2-methyl-3-buten-2-ol (k1), 3-methyl-2-buten-1-ol (k2), 3-methyl-3-buten-1-ol (k3), 2-methyl-3-buten-1-ol (k4), and 3-methyl-3-buten-2-ol (k5) at 298 ± 2 K and 1000 ± 10 mbar total pressure of synthetic air were determined under low- and high-NOx conditions using the relative kinetic technique. The present work provides for the first time the rate coefficients of gas-phase reactions of hydroxyl radicals with 2-methyl-3-buten-1-ol and 3-methyl-3-buten-2-ol. The following rate constants were measured (in 10–11 cm3 molecule–1 s–1): k1 = 6.32 ± 0.49, k2 = 14.55 ± 0.93, k3 = 10.04 ± 0.78, k4 = 5.31 ± 0.37, and k5 = 11.71 ± 1.29. No significant differences in the measured rate coefficients were obtained when either 365 nm photolysis of CH3ONO in the presence of NO or 254 nm photolysis of H2O2 was used as a source of OH radicals. Reactivity toward other classes of related compounds such as alkenes and saturated alcohols is discussed. A comparison of the structure–activity relationship (SAR) estimates derived from the available accepted methodologies with experimental data available for unsaturated alcohols is provided. Atmospheric lifetimes for the investigated series of alkenols with respect to the main atmospheric oxidants are given and discussed.
Introduction
The atmospheric chemistry of biogenic volatile organic compounds (BVOCs) received a large interest in the last decades.1−6 In particular, there is ongoing research in the atmospheric chemistry of methyl-butenols (MBOs) due to their reactivity toward the main atmospheric oxidants (OH radicals, O3, NO3 radicals, or Cl atoms)6−11 and due to their importance in photo-oxidants formation in the lower atmosphere.12
Methyl-butenols in the form of 2-methyl-3-buten-2-ol (MBO232), 3-methyl-2-buten-1-ol (MBO321), 3-methyl-3-buten-1-ol (MBO331), 2-methyl-3-buten-1-ol (MBO231), and 3-methyl-3-buten-2-ol (MBO332) are among atmospheric constituents emitted from biogenic sources, like pine forests,13−16 blossoming rye,17 and deciduous forests.18 The most studied MBO is MBO232 due to its similarities with isoprene, a well-known biogenic compound emitted in large quantities from vegetation.19 Emission fluxes of MBO232 are strongly dependent on the actinic flux during the daylight hours.9 In Northern American pine forests, the level of methyl-butenol is 5–8 times larger than that of isoprene.9,13 MBO emissions seem to be at the minimum during low-temperature periods with decreased light intensity, i.e., in winter.20
The main atmospheric removal path of biogenic compounds is the reaction with photochemical oxidants such as OH radicals, ozone (O3), nitrate radical (NO3), and chlorine atoms (Cl).21 Photolysis is a minor atmospheric sink path for some biogenic alcohols.22,23 The OH radical, called also the atmospheric detergent, reacts in the gas phase with almost all of the volatile organic compounds (VOCs).24 Unsaturated compounds, including also alkenols, interact with the OH radicals via two distinct channels: (1) the OH-radical addition to the unsaturated bonds and (2) the H-atom abstraction from either the C atoms or from the OH alcohol functional group, with the first channel accounting for more than 70%.12 Their atmospheric oxidation process, with respect to their emission areas, leads to the formation of atmospheric peroxy (HO2/RO2) photo-oxidants, secondary organic aerosols (SOA), and carbonyls (i.e., formaldehyde, acetaldehyde, propanaldehyde, acetone). In areas with high NOx concentration, the sink of alkenols can contribute to the peroxyacetyl nitrate formation.25
Experimental kinetic studies have been reported for the reaction of OH radicals with MBO232 by Rudich et al.,26 Fantechi et al.,27 Ferronato et al.,28 Papagni et al.,29 Imamura et al.,30 Baasandorj and Stevens,31 Carrasco et al.,32 and Takahashi et al.33 Available rate constant values vary in the range of (3.90–6.90) × 10–11 cm3 molecule–1 s–1 at 298 ± 2 K. Moreover, the gas-phase oxidation mechanism of MBO232 with OH radicals has been investigated, with acetone, glycolaldehyde, 2-hydroxy-2-methylpropanal (HMPR), and formaldehyde26,28,32−37 reported as reaction products. Rudich et al.26 found that MBO232 reacts 15–20% faster with the OH radicals in the presence of O2. Moreover, a negative dependence of the temperature for the k-values was reported. Also, the OH addition to the double bond seems significantly to occur (∼67%) through the anti-Markovnikov addition at the terminal C(sp2) carbon,25 and this is consistent with a Lindemann–Hinshelwood reaction mechanism where the reaction of O2 with the MBO–OH adduct is the limiting reaction step.26 Experimental and computational investigations performed by Baasandorj and Stevens31 seem to be in good agreement with the experimental observation of Rudich et al.,26 the computational results indicating that the OH radicals would add to the internal C atom through a Markovnikov addition type, nearest to the alcohol group.
Imamura et al.30 carried out studies for the OH-radical-initiated reaction of MBO321 and reported a reaction rate coefficient of (1.50 ± 0.1) × 10–10 cm3 molecule–1 s–1 at 298 K. The gas-phase reactions rate coefficient for MBO331 with OH radicals was investigated using the relative-rate technique30 and PLP-LIF (pulsed laser photolysis-laser-induced fluorescence) absolute method,38 and the rate coefficient of (9.70 ± 0.70) × 10–11, (9.70 ± 1.80) × 10–11, and (9.40 ± 0.40) × 10–11 cm3 molecule–1 s–1, respectively, were reported.
Knap et al.11 used ab initio calculations and transition state theory to investigate the potential H-shift reaction in four MBO peroxy radicals. In environments characterized by high NOx levels, approximately 85% of MBO232-peroxy radical reactions occur with NO. However, in remote areas, the MBO232-peroxy radical sink shifts toward reactions with HO2.
There have been no reported kinetic measurements for the reaction of MBO231 and MBO332 with OH radicals. The present work reports on kinetic data and reaction pathways presumptions for the following methyl-butenols:
![]() |
1 |
![]() |
2 |
![]() |
3 |
![]() |
4 |
![]() |
5 |
The kinetics results from the present study extend the available kinetic database, needed to develop and update the structure–activity relationship methodologies used in global modeling to assess the atmospheric fate of various BVOCs. Five different structural–activity relationship (SAR) approaches proposed in the current literature24,39−41 were used in the present work to estimate the rate coefficients for the investigated reactions. The results are compared and discussed with respect to the available experimental information, SAR estimated values, and atmospheric chemistry of methyl-butenols.
Experimental Section
The measurements were performed in the Environmental Simulation Chamber made of Quartz from the “Alexandru Ioan Cuza” University of Iasi (ESC-Q-UAIC), Romania. Investigations were performed at a temperature of 298 ± 2 K and 1000 ± 10 mbar total pressure of synthetic air. The chamber consists of three quartz glass tubes and is surrounded by 32 Philips TL-DK (36 W) actinic lamps with an emission maximum at 365 nm and 32 Philips UV-C TUV 30W/G30 T8 germicidal lamps with an emission maximum at 254 nm. A multiple reflection White-type mirror system with a total optical path length of 492 ± 1 m is mounted inside the reactor. A Bruker Vertex 80 FTIR spectrometer is coupled to the reflection mirror system. Details concerning the ESC-Q-UAIC chamber construction and facilities and validation approaches are reported in another paper.42 Quantitative analysis of the gas-phase mixture was accomplished by recording infrared spectra over the range 700–4000 cm–1 at a spectral resolution of 1 cm–1. IR spectra were recorded over a time period of 15 to 25 min at a time resolution of 1 min.
Relative-rate technique has been used to determine the gas-phase rate coefficients for OH-initiated reactions of methyl-butenols. Propene, trans-2-butene, 1,3,5-trimethylbenzene, and 1,2-dihydroxybenzene were used as reference compounds.
Prior to the initiation of the OH-radical reactions, the reactor wall-loss rates were monitored for all of the investigated compounds as well as for the reference compounds. Also, all of the compounds were investigated for photolysis losses upon starting the reactions with the OH radicals. The chemicals were used as received, without any further 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), propene (Aldrich, 99,5%), trans-2-butene (Aldrich, 99%), 1,3,5-trimethylbenzene (Aldrich, 98%), 1,2-dihydroxybenzene (Aldrich, 98%), hydrogen peroxide (Aldrich, 40% solution in water), and synthetic air (Messer-Griesheim, 99.999%). Methyl nitrite was prepared and stored as described previously.42
Liquid methyl-butenols and gaseous reference compounds were transferred into the reaction vessel through a preheated inlet line in a flow of synthetic air using syringes and microsyringes. Solid compounds MBO332 and 1,2-dihydroxybenzene were transferred into the chamber at low pressure via a preheated glassware port. Initial concentrations of the analytes were determined based on the weighted amount of the transferred compound. The initial concentration of the investigated methyl-butenols ranged between 3.77 and 11.60 × 1013 molecules cm–3 and those of the reference compounds between 3.18 and 22.70 × 1013 molecules cm–3. Regarding the OH-radical precursors, their concentrations (in molecules cm–3) were in the range of (6.49–19.46) × 1013 for methyl nitrite and (1.64–2.45) × 1014 for hydrogen peroxide.
In the presence of OH radicals, both the methyl-butenols and the reference compounds are oxidized by the following reaction sequence:
![]() |
6 |
![]() |
7 |
The rate coefficient values used in the present study for the reference compounds (in 10–11 cm3 molecule–1 s–1) are as in the following: k(propene + OH) = (2.63 ± 0.39),43k(trans-2-butene + OH) = (6.31 ± 0.95),24k(1,3,5-trimethylbenzene + OH) = (5.86 ± 0.88),21 and k(1,2-dihydroxybenzene + OH) = (10.4 ± 1.6).44
For the wall-loss evaluation, just before the beginning of each set of experiments, the reaction mixtures were allowed to stand in the dark for 10 min and then were subjected to UV irradiation for 10 min to evaluate the photolysis losses of the organic reagents. The losses due to wall deposition or photolysis processes follow a first-order kinetics:
![]() |
8 |
![]() |
9 |
![]() |
10 |
The relative kinetic equation used to experimentally evaluate the OH gas-phase reaction rate constants of MBOs is described by eq I.
![]() |
I |
where [MBO]t0 and [MBO]t are the concentrations of the methyl-butenols at times t0 and t, respectively; [reference]t0 and [reference]t are the concentrations of the reference compound at times t0 and t, respectively; and kMBO and kref are the rate coefficients of the reactions of methyl-butenols (1–5) and reference compounds with OH radicals, respectively. Hence, plots of (ln([reactant]t0/[reactant]t) – (kMBO,WL × (t – t0)) against (ln([reference]t0/[reference]t) – (kref,WL + kref,J) × (t – t0)) should be straight lines with zero intercept and slope of kMBO/kref, from which the rate coefficients for the reactants can be placed on absolute base.
In the present study, two in situ OH-radical sources were used, i.e., 365 nm photolysis of methyl nitrite (CH3ONO)/NO mixture and 254 nm photolysis of hydrogen peroxide (H2O2), and the following reactions accounting for the OH-radical formation:
![]() |
11 |
![]() |
12 |
![]() |
13 |
![]() |
14 |
Structure–Activity Relationship Methodology
Gas-phase rate coefficients of OH-radical reactions with methyl-butenols are essential parameters that allow the assessment of the dominant tropospheric removal processes of unsaturated alcohols.
In the present work, five different SAR approaches were used to evaluate the most suitable method to predict the gas-phase reactivity of this series of methyl-butenols toward OH radicals.
Calvert et al.24 explore SAR methodologies proposed by Atkinson et al.39,45,46 and the method proposed by Peeters et al.47 The SAR concept of Peeters et al. suggests that the estimation of the rate coefficients can be described solely as the sum of the rate coefficients for independent site-specific OH radicals addition at each C atom in the C=C bond. The SAR method from Peeters et al.47 is recommended to be used for alkenes and polyalkenes. The SAR method from Kwok and Atkinson39 is a well-known method for estimating the total reaction rate coefficient as a sum of the reaction pathways involving the H-atom abstraction from C–H and O–H bonds and the OH-radical addition to the double bond. This method considers the double bond as one addition site and treats it as another substituent group. AOPWINv1.92 automated mechanism available in EPI-SUITE software,48 developed by US-EPA, was also used to estimate the reaction rate coefficients of alkenols. This application employs Kwok and Atkinson39 methodology with some updated substituent factors. Pfrang et al.40 proposed an estimation methodology that uses the correlation between experimental kinetic values and the energy of the highest occupied molecular orbital (HOMO) to accommodate group reactivity factors on basic structures of C=C bonds. Pfrang et al.40 method does not account for the H-atom abstraction channels, this route being discretely included in the R substituent factors due to the build of the SAR method. More recently, Jenkin et al.41 developed a SAR method based on updating and optimizing the methods presented above with the latest available kinetic data. The total rate coefficient is defined as a sum of the rate coefficients for H-atom abstraction and OH addition at each C atom involved in the formation of the double bond, combining the principles described by Kwok and Atkinson39 and Peeters et al.47
Results and Discussion
Rate Coefficient Values
The second-order rate coefficients for reactions 1–5 were measured under simulated atmospheric conditions at 298 ± 2 K and 1000 ± 10 mbar total air pressure. The values were derived from the relative loss of methyl-butenols vs the loss of the reference compounds in the presence of OH radicals. Data from both low-NOx and high-NOx measurement conditions were corrected for wall loss and photolysis processes. The loss coefficients (kMBO,WL, kref,WL, and kref,hν) measured in the ESC-Q-UAIC chamber setup are as follows: kMBO321,WL = (1.21 ± 0.07) × 10–4 s–1 for MBO321, kMBO331,WL = (3.75 ± 0.28) × 10–5 s–1 for MBO331, kTMB,WL = (4.22 ± 0.19) × 10–5 s–1 for 1,3,5-trimethylbenzene, and kCAT,WL = (1.51 ± 0.15) × 10–4 s–1 for 1,2-dihydroxybenzene and kTMB,J = (1.40 ± 0.05) × 10–4 s–1 for 1,3,5-trimethylbenzene, with uncertainties corresponding for two standard deviation (2σ). Preliminary tests have shown that wall deposition is specific for MBO321 and MBO331, 1,3,5-trimethylbenzene and 1,2-dihydroxybenzene. Upon irradiation at 365 or 254 nm, no additional significant losses were observed either for MBOs or for reference compounds, except for 1,3,5-trimethylbenzene, which photodissociates under the experimental conditions at 254 nm only.
Variables in eq I were plotted separately, and linear regression analysis was applied to each individual data set in order to determine the kMBO/kref ratios. Figures 1–5 show the kinetic plots derived from the experimental data. Straight-line distributions can be observed in all cases. Relative rate ratios (kMBO/kref) and reaction rate coefficients for all of the investigated reactions are presented in Table 1. Uncertainties for the kMBO/kref represent 2σ from the linear regression analysis. The kMBO rate coefficient uncertainties (Δk) are given from a combination of 2σ values from the linear regression and the uncertainty from the selected reference compound as expressed in eq II.
![]() |
II |
![]() |
III |
![]() |
IV |
The k(average) and its uncertainty (Δk(average)) were calculated using the weighted average of the k(MBO + OH) values, where the weight (ωi) was calculated using eq III and the uncertainty was calculated using eq IV. As observed in Figures 1–5, no significant differences have been obtained under the present experimental conditions for the MBOs rate coefficients when both low or high NOx concentrations were involved. Thus, the rate coefficient presented for each studied MBO represents the weighted average of the kMBO(average) values determined when different OH-radical sources were used.
Figure 1.
Relative loss of MBO232 vs the reference compounds, (○/●) trans-2-butene and (Δ/▲) 1,3,5-trimethylbenzene (empty symbols belong to low-NOx conditions and filled symbols are from high-NOx conditions). The solid regression line corresponds to MBO232 vs trans-2-butene and medium dashed regression line to MBO232 vs 1,3,5-trimethylbenzene.
Figure 5.
Relative loss of MBO332 vs the reference compounds, (●) trans-2-butene and (×) 1,2-dihydroxybenzene in high-NOx conditions. The solid regression line corresponds to MBO332 vs trans-2-butene and short dashed regression line to MBO332 vs 1,2-dihydroxybenzene.
Table 1. Gas-Phase Rate Coefficient Values of OH-Radical Reactions with All Investigated Methyl-Butenols in the Present Study.
methyl-butenol | OHsource | reference compounds | k/kref | k × 1011 (cm3 molecule–1 s–1) | k(average) × 1011 (cm3 molecule–1 s–1) | k(total) × 1011 (cm3 molecule–1 s–1) |
---|---|---|---|---|---|---|
2-methyl-3-buten-2-ol (MBO232) | CH3ONO | trans-2-butene | 1.02 ± 0.02 | 6.44 ± 0.98 | 6.48 ± 0.70 | 6.32 ± 0.49 |
1,3,5-trimethylbenzene | 1.11 ± 0.04 | 6.52 ± 1.01 | ||||
H2O2 | trans-2-butene | 0.99 ± 0.04 | 6.31 ± 0.99 | 6.18 ± 0.68 | ||
1,3,5-trimethylbenzene | 1.04 ± 0.03 | 6.07 ± 0.68 | ||||
3-methyl-2-buten-1-ol (MBO321) | CH3ONO | propene | 5.76 ± 0.22 | 15.14 ± 2.32 | 14.86 ± 1.17 | 14.55 ± 0.93 |
1,2-dihydroxybenzene | 1.72 ± 0.05 | 17.84 ± 2.93 | ||||
1,3,5-trimethylbenzene | 2.18 ± 0.09 | 12.76 ± 1.98 | ||||
trans-2-butene | 2.48 ± 0.09 | 15.65 ± 2.43 | ||||
H2O2 | trans-2-butene | 2.15 ± 0.04 | 13.54 ± 2.05 | 14.04 ± 1.51 | ||
1,3,5-trimethylbenzene | 2.50 ± 0.08 | 14.64 ± 2.24 | ||||
3-methyl-3-buten-1-ol (MBO331) | CH3ONO | trans-2-butene | 1.58 ± 0.04 | 9.99 ± 1.52 | 10.05 ± 1.10 | 10.04 ± 0.78 |
1,2-dihydroxybenzene | 0.97 ± 0.03 | 10.11 ± 1.58 | ||||
H2O2 | trans-2-butene | 1.56 ± 0.07 | 9.85 ± 1.55 | 10.04 ± 1.10 | ||
1,3,5-trimethylbenzene | 1.75 ± 0.04 | 10.23 ± 1.55 | ||||
2-methyl-3-buten-1-ol (MBO231) | CH3ONO | trans-2-butene | 0.86 ± 0.02 | 5.40 ± 0.83 | 5.46 ± 0.48 | 5.31 ± 0.37 |
1,2-dihydroxybenzene | 0.52 ± 0.02 | 5.37 ± 0.84 | ||||
propene | 2.13 ± 0.07 | 5.61 ± 0.85 | ||||
H2O2 | trans-2-butene | 0.83 ± 0.03 | 5.25 ± 0.81 | 5.13 ± 0.56 | ||
1,3,5-trimethylbenzene | 0.86 ± 0.02 | 5.02 ± 0.76 | ||||
3-methyl-3-buten-2-ol (MBO332) | CH3ONO | trans-2-butene | 1.75 ± 0.06 | 11.07 ± 1.71 | 11.71 ± 1.29 | 11.71 ± 1.29 |
1,2-dihydroxybenzene | 1.21 ± 0.03 | 12.55 ± 1.96 |
Figure 2.
Relative loss of MBO321 vs the reference compounds, (○/●) trans-2-butene, (Δ/▲) 1,3,5-trimethylbenzene, (■) propene, and (×) 1,2-dihydroxybenzene in high-NOx conditions (empty symbols belong to low-NOx conditions and filled symbols are from high-NOx conditions). The solid regression line corresponds to MBO321 vs trans-2-butene, medium dashed regression line to MBO321 vs 1,3,5-trimethylbenzene, dotted regression line to MBO321 vs propene, and short dashed regression line to MBO321 vs 1,2-dihydroxybenzene.
Figure 3.
Relative loss of MBO331 vs the reference compounds, (○/●) trans-2-butene, (Δ) 1,3,5-trimethylbenzene, and (×) 1,2-dihydroxybenzene in high-NOx conditions (empty symbols belongs to low-NOx conditions and filled symbols are from high-NOx conditions). The solid regression line corresponds to MBO331 vs trans-2-butene, medium dashed regression line to MBO331 vs 1,3,5-trimethylbenzene, and short dashed regression line to MBO331 vs 1,2-dihydroxybenzene.
Figure 4.
Relative loss of MBO231 vs the reference compounds, (○/●) trans-2-butene, (Δ) 1,3,5-trimethylbenzene, (■) propene, and (×) 1,2-dihydroxybenzene in high-NOx conditions (empty symbols belong to low-NOx conditions and filled symbols are from high-NOx conditions). The solid regression line corresponds to MBO231 vs trans-2-butene, medium dashed regression line (overlapped over the solid line) to MBO231 vs 1,3,5-trimethylbenzene, dotted regression line to MBO231 vs propene, and short dashed regression line to MBO231 vs 1,2-dihydroxybenzene.
Table 2 presents the gas-phase rate coefficients of OH radicals with methyl-butenols as determined in the present work among other available literature values at a temperature of 298 K and in the presence of O2. Control tests were performed during the kinetic investigations, monitoring the OH reaction rate coefficient ratios for the reference compounds. Figure S1A,B shows the linear distribution of the experimental kinetic data for the references under both high- and low-NOx conditions, respectively. Table S1 presents the linear regression analysis for the control tests performed during the MBO kinetic measurements. Deviations less than 15% from the reference values used to place the MBOs reaction rate coefficients on an absolute scale were obtained, which is in very good agreement with currently used kinetic databases.21
Table 2. Rate Coefficient Values for the Gas-Phase Reaction between Various Methyl-Butenols and OH Radicals, as Determined in the Present Work or Available from the Literature.
methyl-butenol | k × 1011 (cm3 molecule–1 s–1) | methoda | reference |
---|---|---|---|
2-methyl-3-buten-2-ol (MBO232) | 6.32 ± 0.49 | RR | this work |
6.40 ± 0.60 | AR | Rudich et al.26 | |
3.90 ± 1.20 | RR | Fantechi et al.27 | |
6.90 ± 1.00 | RR | Ferronato et al.28 | |
5.67 ± 0.13 | RR | Papagni et al.29 | |
6.60 ± 0.50 | RR | Imamura et al.30 | |
6.61 ± 0.66 | AR | Baasandorj and Stevens31 | |
5.60 ± 0.60 | RR | Carrasco et al.32 | |
6.49 ± 0.82 | RR | Takahashi et al.33 | |
3-methyl-2-buten-1-ol (MBO321) | 14.55 ± 0.93 | RR | this work |
15.00 ± 1.00 | RR | Imamura et al.30 | |
3-methyl-3-buten-1-ol (MBO331) | 10.04 ± 0.78 | RR | this work |
9.70 ± 0.70 | RR | Imamura et al.30 | |
9.70 ± 1.80 | RR | Cometto et al.38 | |
9.40 ± 0.40 | AR | Cometto et al.38 | |
2-methyl-3-buten-1-ol (MBO231) | 5.31 ± 0.37 | RR | this work |
3-methyl-3-buten-2-ol (MBO332) | 11.71 ± 1.29 | RR | this work |
AR: the absolute rate method; RR: the relative-rate method.
The k1 value obtained in the present work is in very good agreement with the values previously reported by Ferronato et al.,28 Rudich et al.,26 Takahashi et al.,33 Imamura et al.,30 Baasandorj and Stevens,31 and Carrasco et al.32 However, for MBO232, the k1 value is slightly higher than that proposed by Papagni et al.29 using the relative-rate method and 1,3,5-trimethylbenzene as the reference compound. The value reported by Fantechi et al.27 of (3.90 ± 1.20) × 10–11 cm3 molecule–1 s–1 is about 40% smaller than the value proposed in this study. Fantechi et al.27 carried out relative-rate measurements for reaction 1 using the in situ photolysis of hydrogen peroxide as a source of OH radicals and isoprene and propene as the reference compounds.
The reaction rate coefficient available for MBO321 (k2) was experimentally determined at 298 K and 760 Torr of air by Imamura et al.,30 and the value of (15.00 ± 1.00) × 10–11 cm3 molecule–1 s–1 is in good agreement with the current reported value of (14.55 ± 0.93) × 10–11 cm3 molecule–1 s–1. The reaction rate coefficient (k3) for OH + MBO331 was measured at 298 K by Imamura et al.30 and Cometto et al.38 using the relative-rate technique. Their reported values of (9.70 ± 0.70) × 10–11 and (9.70 ± 1.80) × 10–11 cm3 molecule–1 s–1 are in good agreement with that of (10.04 ± 0.78) × 10–11 cm3 molecule–1 s–1 evaluated within the present work. Also, pulsed laser photolysis-laser-induced fluorescence (PLP-LIF) measurements were undertaken by Cometto et al.38 to determine the OH reaction rate coefficient for MBO331 at 107 Torr of helium. The resulting reaction rate coefficient value of (9.40 ± 0.40) × 10–11 cm3 molecule–1 s–1 is approximately 7% lower than the value reported in the present study. For MBO231 and MBO332, the OH-reaction rate coefficients obtained in this study are k4 = (5.31 ± 0.37) × 10–11 cm3 molecule–1 s–1 and k5 = (11.71 ± 1.29) × 10–11 cm3 molecule–1 s–1. No other reaction rate coefficients have been reported for the reactions of the OH radical with MBO231 and MBO332.
Experimental results show that the reactivity toward OH radicals in the gas phase follows the order: k(MBO321) > k(MBO332) > k(MBO331) > k(MBO232) > k(MBO231). The results are consistent when accounting for the stability of the MBO–OH adducts generated from the addition of the OH radical to the double bond.26,31,38 Based on the literature data, Figure 6 presents the possible adducts formed from the OH radical attack on the investigated MBOs. Unlike other MBOs in the series, MBO321-OH adduct stability is ensured by the formation of Ctert(•) or Csec(•) radicals, the first one generated through the OH addition to the nearest C(sp2)-atom from the alcohol group and the second by the OH addition to the further C(sp2) atom from the alcohol group. For MBO232 and MBO231, their adducts contain Csec(•) and Cprim(•); thus, their stability and also their reactivity toward the OH radicals are the least in the series. Differences in the reactivity toward the OH radicals of MBO321 compared with MBO331 and MBO332 and that of MBO232 and MBO231 show clearly that the OH addition occurs mainly to the least substituted C atom in the double bond. In these situations, the steric hindrance could also play a role in the formation of the most stable OH adduct by diminishing the OH-radical addition to the most substituted C atom, forming the double bond, as seen also in the alkene series.49
Figure 6.
Possible MBO adducts from OH-radical interaction with the investigated.
Data acquired under atmospheric conditions suggest no dependence of MBO OH-reaction rate coefficients on the NOx levels. This is consistent with the available literature data for NOx and NOx-free conditions. For MBO232, an O2 pressure dependence was experimentally observed by Rudich et al.26 and Baasandorj and Stevens,31 which is consistent with the implications for the OH addition to the C=C bond. From the kinetic findings, an attempt at the degradation mechanism of MBOs suggests that the presence of NOx could facilitate the formation of the β,γ-dihydroxy alkoxy radicals and, subsequently, the formation of formaldehyde and 2-hydroxy-2-methylpropanal with a molar yield of approximatively 31% from the oxidation of MBO232 via a minor Marovnikov OH-addition route.32,33
Comparing the Experimental Data through Structure–Activity Relationship (SAR) Methodologies
Table 3 presents, for comparison purposes, the experimental k values (kMBO) as determined in the present work together with the values estimated with different SAR approaches (kSAR). Tables S2–S6 show the procedure of kSAR calculations for each MBO using the SAR methodologies.
Table 3. Rate Coefficient Values for the Reaction of OH Radicals with Various Methyl-Butenols in Direct Relation to with Different Structural–Activity Relationship (SAR) Methods.
Following the least-square fit analysis for experimentally observed (kMBO) and calculated (kSAR) reaction rate coefficients for the five MBOs, SAR estimates from Jenkin et al.41 show the best fit with the experimental (∑((kMBO – kSAR)/kMBO)2 = 0.124), within 25% of the observed rate coefficients. Also, Pfrang et al.40 method appears as the second best fit (∑((kMBO – kSAR)/kMBO)2 = 0.284), followed by Kwok and Atkinson39 (∑((kMBO – kSAR)/kMBO)2 = 0.511), Peeters et al.47 (∑((kMBO – kSAR)/kMBO)2 = 0.927), and AOPWINv1.9248 (∑((kMBO – kSAR)/kMBO)2 = 1.060). Pfrang et al.40 SAR provides reliable estimates, but the method cannot be extended for other classes of compounds. Kwok and Atkinson39 SAR estimates for alkenols are lower than the experimentally observed ones. Large differences are observed when using AOPWINv1.9248 in comparison with Kwok and Atkinson39 methodology for three out of five studied MBOs. SAR estimated values from the considered methodologies for a series of 21 unsaturated alcohols are presented in Table S7. For 15 out of 21 estimates the AOPWINv1.9248 subestimates the reaction rate coefficients almost by a factor of 1.6 compared with the initial methodology of Kwok and Atkinson.39 For the OH-addition channel AOPWINv1.9248 does not account for the C(−CH2–OH) or any C(–CR2–OH) (R = alkyl) substituent as proposed in the initial method.39 Differences could also be noticed for the H-atom abstraction channel where a later update of AOPWIN introduced a new substituent factor F(>C–OH) = 3.80 derived from experimental data obtained for tertiary alcohols (i.e., MBO232 and 3-methyl-1-penten-3-ol).
However, differences between estimates could reflect different approaches in the SAR methodologies. Pfrang et al.35 evaluate the rate coefficients of the basic alkenols double bond, differently for those for alkenes, and propose scaling substituent factors, f(R), which are dependent on the substituent length, but neglecting the H-atom abstraction channels. Kwok and Atkinson34 consider in their approach the number, identity, and position of the substituent groups around the R1R2C=CR3R4 structural unit as the main factors influencing the addition of the OH radicals. In the Jenkin et al.41 approach, the addition of the OH radicals is estimated for each C atom involved in the formation of the double bond, and it is influenced by the number and type of the substituents. Both Kwok and Atkinson39 and Jenkin et al.41 methodologies evaluate the rate coefficients corresponding to the H-atom abstractions channels. However, the H-atom abstraction either from the OH-group or from alkyl-groups has a minor effect compared to the OH addition channel for MBOs (Table S7). For higher C-chain alkenols (C6–C8), the contribution from the H-atom abstraction channel over the overall reaction rate coefficient gains more weight in the Jenkin et al.41 method compared with that of Kwok and Atkinson,39 in good agreement with experimentally observed kinetic data presented in this work.
Correlation between OH Reaction Rate Coefficients for Alkenols and Other Homologue Compounds
The addition of the OH radicals to the double bond in unsaturated alcohols form β-hydroxyalkyl radicals,35,39,50 which seems to be the main reaction pathway in the gas phase. Table 4 presents, for comparison purposes, the reactivity of alkenols sharing structural similarities with the investigated MBOs and their corresponding parent alkene and saturated alcohols toward the OH radicals. As revealed by the data in Table 4, the presence of the hydroxyl substituent makes a significant difference, with the OH reaction rate coefficients of the alkenols being almost twice those for the alkenes. An enhanced reactivity can also be observed while comparing the rate coefficients of the MBOs with those of their homologue-saturated alcohols. Data in Table 4 clearly show that the alkenols react 5 up to 18 times faster with the OH radicals than the saturated alcohols. The unsaturation of the carbon chain provides a high electron density cloud and opens the OH-radical addition channel to the double bond, hence increasing the reactivity of the interest MBOs more than 1 order of magnitude (e.g., MBO321, MBO332, and MBO232).
Table 4. Rate Coefficients (in Units of cm3 Molecule–1 s–1) for the Gas-Phase Reaction of OH Radicals with Alkenols vs Their Alkene Structural Homologues and Alkenols vs Their Saturated Alcohols Structural Homologues at a Temperature of 298 ± 4 K.
alkenols | kOH × 1011 | alkenes | kOH × 1011 | kalkenol/kalkene | alcohols | kOH × 1011 | kalkenol/kalcohol |
---|---|---|---|---|---|---|---|
2-propen-1-ol | 5.0021 | propene | 2.4421 | 2.05 | propan-1-ol | 0.5921 | 8.47 |
2-methyl-2-propen-1-ol | 9.0021 | 2-methylpropene | 5.1021 | 1.82 | 2-methylpropan-1-ol | 0.9721 | 9.59 |
3-buten-1-ol | 5.5021 | 1-butene | 3.1454 | 1.75 | butan-1-ol | 0.9121 | 6.04 |
2-buten-1-ol | 8.7021 | (Z/E)2-butene | 5.28/6.3121 | 1.65/1.38 | 9.56 | ||
3-methyl-2-buten-1-ol | 14.55a | 2-methyl-2-butene | 8.6945 | 1.67 | 3-methyl-butan-1-ol | 1.3955 | 10.47 |
3-methyl-3-buten-1-ol | 10.04a | 2-methyl-1-butene | 6.1043 | 1.65 | 7.22 | ||
3-methyl-3-buten-2-ol | 11.71a | 1.92 | 3-methyl-butan-2-ol | 1.2555 | 9.37 | ||
2-methyl-3-buten-2-ol | 6.23a | 3-methyl-1-butene | 3.1845 | 1.99 | 2-methyl-butan-2-ol | 0.3622 | 17.56 |
2-methyl-3-buten-1-ol | 5.31a | 1.67 | 2-methyl-butan-1-ol | ||||
(Z)2-penten-1-ol | 11.7021 | Z-2-pentene | 6.5021 | 1.80 | pentan-1-ol | 1.1021 | 10.64 |
(E)2-penten-1-ol | 6.7656 | E-2-pentene | 6.7021 | 1.01 | 6.15 | ||
1-penten-3-ol | 6.3321 | 1-pentene | 3.2221 | 1.97 | pentan-3-ol | 1.3021 | 4.87 |
(Z)2-hexen-1-ol | 9.7721 | hexan-1-ol | 1.3021 | ||||
(E)2-hexen-1-ol | 7.9321 | (E)2-hexene | 6.0857 | 1.30 | 6.10 | ||
(Z)3-hexen-1-ol | 10.8021 | ||||||
(E)3-hexen-1-ol | 9.7321 | (Z)3-hexene | 6.2758 | 1.72 | 8.31 | ||
(Z)4-hexen-1-ol | 8.0821 | ||||||
(E)4-hexen-1-ol | 7.1421 | ||||||
3-methyl-1-penten-3-ol | 6.2059 | 3-methylpentan-1-ol | |||||
(Z)3-hepten-1-ol | 12.8060 | heptan-1-ol | 1.3021 | 9.85 | |||
(Z)3-octen-1-ol | 14.9060 | octan-1-ol | 1.3021 | 11.46 |
This work.
Oxygen atom from the alcohol functional group polarizes both the C–O bond and the O–H bond in saturated or unsaturated alcohols. The presence of unsaturation in the compound leads to an increase in the polarizability of some bonds. Data presented in Table 4 show that the values of the rate constants ratios, kalkenol/kalcohol, differ considerably by a factor of 4.8–17.7. It is known that the reactivity of saturated alcohols increases from primary alcohol > secondary alcohol > tertiary alcohol,12,51,52 and this holds for 3-methyl-butan-1-ol > 3-methyl-butan-2-ol > 2-methyl-butan-2-ol. The alcohol functional group stabilizes the hydroxyalkyl radical generated from the H-atom abstraction from β- or γ C atom; thus, for higher C-chain alcohols, the reactivity does not increase anymore.
In the case of unsaturated alcohols, the order of reactivity is dependent upon the stability of the OH-adduct, with Ctert(•) being more stable than Csec(•), which is more stable than Cprim(•), as discussed earlier. The influence of the alcohol group is not so evident in the first instance. Considering the most stable adduct formation, the presence of two hydroxyl groups near the radical site could stabilize the nearly formed adducts. The β,β′-dihydroxy radical formed from MBO232 gains supplementary stability from interactions with the OH groups in comparison with its isomer MBO231, which generates the most stable adduct in the form of an β,γ-dihydroxy radical. Similar observations are also valid for MBO332 and MBO331. This supplementary stabilization is an indication of the 15–17% difference between the reactivity of allylic alcohols and γ-alkenols. The presence of the alcohol functional group near the double bond increases the reactivity in comparison with the parent alkenes with a factor of about 1.7, consistent with the substituent factor for the OH-addition channel F′(CH2–OH) or similar of 1.8, proposed by Jenkin et al.41
Considerations on the Atmospheric Implications of the OH-Radical-Initiated Oxidation of Various MBOs
Table 5 presents the tropospheric lifetimes for the investigated alcohols. The experimental rate constants obtained in the present work were used to estimate the tropospheric lifetime of the five biogenic unsaturated alcohols toward OH radicals. To identify the main atmospheric removal pathway of the investigated methyl-butenols, their tropospheric lifetimes were determined in comparison to their reactions with other major atmospheric oxidants, i.e., O3, NO3 radicals, and Cl atoms. The atmospheric lifetimes were obtained using the expression: τ = 1/(k[X]), where [X] is the typical atmospheric concentration of the oxidant (OH, O3, NO3, and Cl) and k is the rate coefficient of the reaction of the oxidant with the considered alcohol. The τtotal in Table 5 is obtained using the expression: τtotal = 1/(∑(k[X])).
Table 5. Tropospheric Lifetimes, τ (in Hours), for the Investigated Methyl-Butenols with OH, O3, NO3, and Cl Gaseous Oxidant Speciesa.
rate coefficient (cm3 × molecule–1 × s–1) |
tropospheric lifetime (h) |
||||||||
---|---|---|---|---|---|---|---|---|---|
kOH × 1011b | kO3 × 1017 | kNO3 × 1013 | kCl × 1010 | τOH | τO3 | τNO3 | τCl | τtotal | |
2-methyl-3-buten-2-ol | 6.32 | 0.8319 | 0.1161 | 2.6462 | 3.89 | 47.81 | 50.51 | 105.22 | 3.21 |
3-methyl-2-buten-1-ol | 14.55 | 43.9063 | 10.0061 | 4.0264 | 1.69 | 0.90 | 0.56 | 69.10 | 0.32 |
3-methyl-3-buten-1-ol | 10.04 | 0.6865 | 2.7061 | 4.1365 | 2.45 | 58.36 | 2.06 | 67.26 | 1.10 |
2-methyl-3-buten-1-ol | 5.31 | 0.3765 | (−) | 3.5165 | 4.63 | 106.10 | (−) | 79.14 | <4.17 |
3-methyl-3-buten-2-ol | 11.71 | (−) | (−) | (−) | 2.10 | (−) | (−) | (−) | <2.10 |
In the troposphere, as observed from Table 5, the dominant chemical loss process for methyl-butenols is their daytime reaction with OH radicals since lifetimes are in the range of a few hours. Reaction with NO3 radicals at night comes in the next, while the reactions with O3 or Cl atoms make an insignificant contribution to the tropospheric degradation pathways of MBOs. There is one single MBO, in the form of MBO321, for which the daytime reaction with ozone may be a major pathway loss in the atmosphere. However, one should always keep in mind that the atmospheric lifetime for ozone is 0.5 times higher than the atmospheric lifetime for OH radicals. The loss due to Cl atoms indicates long tropospheric lifetimes at Cl concentrations of 104 molecules cm–3 and could only be of some importance under high Cl concentrations. However, if Cl local peak concentration is as high as 105 molecules cm–3 (Riedel et al.53), then lifetimes turn from 105.22 to 10.52 h for MBO232, 69.10 to 6.91 h for MBO321, 67.26 to 6.73 h for MBO331, and 79.14 h to 7.91 for MBO231. Overall, the studied compounds have high reactivity in the troposphere, and as shown by the estimated tropospheric lifetimes, a significant part of MBOs will be removed in the gas phase in the near vicinity of their emission sources.
Conclusions
Rate coefficient values for the OH-radical reaction with MBO231 and MBO332 are reported for the first time in the present study. The rate coefficients for MBO232, MBO231, and MBO331, as determined in the present work, are in good agreement with previously reported values from other studies. The OH-reactivity trend of MBOs and other alkenols is placed on the basis of the formation of the most stable OH adduct. The alcohol functional group proximity to the double bond may play a role in the further stabilization of the OH adducts generated from the alkenol interaction with the OH radicals.
Five SAR methodologies were used to estimate values of the MBO rate coefficients with OH radicals. The SAR from Jenkin et al. estimated rate coefficients that are closest to those measured experimentally. The reactivity of unsaturated alcohols in comparison with their homologous alkenes series is on average a factor of 1.7 times greater due to the presence of the CH2–OH or (CR2–OH) adjacent to double bond. In comparison with the analogous saturated alcohol series, the reactivity of the investigated MBOs seems to be enhanced mainly due to the presence of the unsaturated bond in the molecule. The use of the estimating AOPWINv1.92 program in further global modeling programs needs to be continuously checked by comparing with the estimates generated according to the initial methodology, as AOPWIN may underestimate the gas-phase OH reaction rate coefficients by 60%.
Tropospheric lifetimes estimated for reactions with the main atmospheric oxidants show that the reaction with OH radicals is the dominant atmospheric fate of MBOs and that they are removed rapidly near their emission sources.
Further investigations of the gas-phase OH oxidation products of higher-chain alkenols are needed to reinforce the current conclusions and to continue the development of the MCM chemical mechanism and SAR models.
Acknowledgments
The authors acknowledge support from the PN-III-P4-PCE2021-0673 (ATMO-SOS) and PN-III-P2-2.1-PED2021-4119 (SOA-REACTOR) UEFISCDI project. 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.4c02287.
Table S1: Kinetic data corresponding to the quality control tests performed during the experimental sessions for MBOs. Table S2: Estimates for the gas-phase reaction of selected methyl-butenols at 298 K calculated according to Pfrang et al. structure–activity relationship (SAR) methodology for alkenols. Table S3: Estimates for the gas-phase reaction of selected methyl-butenols at 298 K calculated according to Peeters et al. structure–activity relationship (SAR) methodology. Table S4: Estimates for the gas-phase reaction of selected methyl-butenols at 298 K calculated according to Kwok and Atkinson structure–activity relationship (SAR) methodology. Table S5: Estimates for the gas-phase reaction of selected methyl-butenols at 298 K from AOPWINv1.92 EPI-SUITE software. Table S6: Estimates for the gas-phase reaction of selected methyl-butenols at 298 K calculated according to Jenkin et al. structure–activity relationship (SAR) methodology. Table S7: OH reaction rate estimates at 298 K for mono-unsaturated alcohols. Figure S1: Representation of the kinetic quality control tests performed during the experimental sessions in low- and high-NOx conditions for MBOs (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- 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]
- Rolletter M.; Kaminski M.; Acir I.-H.; Bohn B.; Dorn H.-P.; Li X.; Lutz A.; Nehr S.; Rohrer F.; Tillmann R.; et al. Investigation of the A-Pinene Photooxidation by OH in the Atmospheric Simulation Chamber SAPHIR. Atmos. Chem. Phys. 2019, 19 (18), 11635–11649. 10.5194/acp-19-11635-2019. [DOI] [Google Scholar]
- Praplan A. P.; Tykkä T.; Schallhart S.; Tarvainen V.; Bäck J.; Hellén H. OH Reactivity from the Emissions of Different Tree Species: Investigating the Missing Reactivity in a Boreal Forest. Biogeosciences 2020, 17 (18), 4681–4705. 10.5194/bg-17-4681-2020. [DOI] [Google Scholar]
- 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 (3), 256. 10.3390/ATMOS11030256. [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, 117306 10.1016/j.atmosenv.2020.117306. [DOI] [Google Scholar]
- Almatarneh M. H.; Elayan I. A.; Abu-Saleh A. A. 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]
- 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]
- Lokachari N.; Wagnon S. W.; Kukkadapu G.; Pitz W. J.; Curran H. J. Experimental and Kinetic Modeling Study of 3-Methyl-2-Butenol (Prenol) Oxidation. Energy Fuels 2021, 35 (17), 13999–14009. 10.1021/acs.energyfuels.1c01530. [DOI] [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]
- Biswas P. Reaction Mechanism and Kinetic Study of the • OH Initiated Tropospheric Oxidation of 3-Methyl-2-Buten-1-Ol: A Quantum Chemical Investigation. J. Theor. Comput. Chem. 2014, 13 (06), 1450052 10.1142/S0219633614500527. [DOI] [Google Scholar]
- 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. 10.1016/j.atmosenv.2016.09.064. [DOI] [Google Scholar]
- Calvert J. G.; Mellouki A.; Orlando J. J.; Pilling M. J.; Wallington T. J.. Mechanisms of Atmospheric Oxidation of the Oxygenates; Oxford University Press, 2011. [Google Scholar]
- 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 (11), 1039–1042. 10.1029/93GL00247. [DOI] [Google Scholar]
- 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. 1998, 103 (D19), 25479–25486. 10.1029/98JD00820. [DOI] [Google Scholar]
- 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. 1999, 104 (D21), 26107–26114. 10.1029/1999JD900749. [DOI] [Google Scholar]
- Schade G. W.; Goldstein A. H. Fluxes of Oxygenated Volatile Organic Compounds from a Ponderosa Pine Plantation. J. Geophys. Res. 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, 414508. 10.1155/2012/414508. [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]
- Koppmann R.Volatile Organic Compounds in the Atmosphere; Koppmann R., Ed.; Wiley: Oxford, UK, 2007. [Google Scholar]
- McGillen M. R.; Carter W. P. L.; Mellouki A.; Orlando J. J.; Picquet-Varrault B.; Wallington T. J. 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]
- Jiménez E.; Lanza B.; Garzón A.; Ballesteros B.; Albaladejo J. Atmospheric Degradation of 2-Butanol, 2-Methyl-2-Butanol, and 2,3-Dimethyl-2-Butanol: OH Kinetics and UV Absorption Cross Sections. J. Phys. Chem. A 2005, 109 (48), 10903–10909. 10.1021/jp054094g. [DOI] [PubMed] [Google Scholar]
- 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 (6), 1831–1837. 10.1021/es8027814. [DOI] [PubMed] [Google Scholar]
- Calvert J. G.; Orlando J. J.; Stockwell W. R.; Wallington T. J.. The Mechanisms of Reactions Influencing Atmospheric Ozone; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
- Saunders S. M.; Jenkin M. E.; Derwent R. G.; Pilling M. J. Protocol for the Development of the Master Chemical Mechanism, MCM v3 (Part A): Tropospheric Degradation of Non-Aromatic Volatile Organic Compounds. Atmos. Chem. Phys. 2003, 3 (1), 161–180. 10.5194/acp-3-161-2003. [DOI] [Google Scholar]
- Rudich Y.; Talukdar R.; Burkholder J. B.; Ravishankara A. R. Reaction of Methylbutenol with the OH Radical: Mechanism and Atmospheric Implications. J. Phys. Chem. A 1995, 99 (32), 12188–12194. 10.1021/j100032a021. [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]
- 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. 1998, 103 (D19), 25,579–25,586. 10.1029/98JD00528. [DOI] [Google Scholar]
- 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 (2), 142–147. . [DOI] [Google Scholar]
- 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 (7), 379–385. 10.1002/kin.20008. [DOI] [Google Scholar]
- 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 (4), 640–649. 10.1021/jp066286x. [DOI] [PubMed] [Google Scholar]
- 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. 2006, 56 (1), 33–55. 10.1007/s10874-006-9041-y. [DOI] [Google Scholar]
- Takahashi K.; Hurley M. D.; Wallington T. J. Kinetics and Mechanisms of OH-initiated Oxidation of Small Unsaturated Alcohols. Int. J. Chem. Kinet. 2010, 42 (3), 151–158. 10.1002/kin.20475. [DOI] [Google Scholar]
- 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 (20), 3547–3556. 10.1016/S1352-2310(98)00061-2. [DOI] [Google Scholar]
- Alvarado A.; Tuazon E. C.; M Aschmann S.; 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 (18), 2893–2905. 10.1016/S1352-2310(99)00106-5. [DOI] [Google Scholar]
- Spaulding R.; Charles M. J.; Tuazon E. C.; Lashley M. Ion Trap Mass Spectrometry Affords Advances in the Analytical and Atmospheric Chemistry of 2-Hydroxy-2-Methylpropanal, a Proposed Photooxidation Product of 2-Methyl-3-Buten-2-Ol. J. Am. Soc. Mass Spectrom. 2002, 13 (5), 530–542. 10.1016/S1044-0305(02)00354-9. [DOI] [PubMed] [Google Scholar]
- 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 (20), 4664–4671. 10.1021/es034142f. [DOI] [PubMed] [Google Scholar]
- Cometto P. M.; Dalmasso P. R.; Taccone R. A.; Lane S. I.; Oussar F.; Daële V.; Mellouki A.; Le Bras G. Rate Coefficients for the Reaction of OH with a Series of Unsaturated Alcohols between 263 and 371 K. J. Phys. Chem. A 2008, 112 (19), 4444–4450. 10.1021/jp7111186. [DOI] [PubMed] [Google Scholar]
- 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 (14), 1685–1695. 10.1016/1352-2310(95)00069-B. [DOI] [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]
- Jenkin M. E.; Valorso R.; Aumont B.; Rickard A. R.; Wallington T. J. Estimation of Rate Coefficients and Branching Ratios for Gas-Phase Reactions of OH with Aliphatic Organic Compounds for Use in Automated Mechanism Construction. Atmos. Chem. Phys. 2018, 18 (13), 9297–9328. 10.5194/acp-18-9297-2018. [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]
- Atkinson R.; Arey J. Atmospheric Degradation of Volatile Organic Compounds Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103 (3), 4605–4638. 10.1021/cr0206420. [DOI] [PubMed] [Google Scholar]
- Olariu R. I.; Barnes I.; Becker K. H.; Klotz B. Rate Coefficients for the Gas-Phase Reaction of OH Radicals with Selected Dihydroxybenzenes and Benzoquinones. Int. J. Chem. Kinet. 2000, 32 (11), 696–702. . [DOI] [Google Scholar]
- Atkinson R. Kinetics and Mechanisms of the Gas-Phase Reactions of the Hydroxyl Radical with Organic Compounds under Atmospheric Conditions. Chem. Rev. 1986, 86 (1), 69–201. 10.1021/cr00071a004. [DOI] [Google Scholar]
- 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 (9), 799–828. 10.1002/kin.550190903. [DOI] [Google Scholar]
- 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 (9), 1618–1631. 10.1021/jp066973o. [DOI] [PubMed] [Google Scholar]
- US EPA . Estimation Programs Interface Suite for Microsoft Windows, v 4.11. https://www.epa.gov/ (accessed April 06, 2023).
- 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]
- Atkinson R.; Aschmann S. M. Rate Constants for the Reactions of O3 and OH Radicals with a Series of Alkynes. Int. J. Chem. Kinet. 1984, 16 (3), 259–268. 10.1002/kin.550160308. [DOI] [Google Scholar]
- McGillen M. R.; Baasandorj M.; Burkholder J. B. Gas-Phase Rate Coefficients for the OH + n -, i -, s -, and t -Butanol Reactions Measured Between 220 and 380 K: Non-Arrhenius Behavior and Site-Specific Reactivity. J. Phys. Chem. A 2013, 117 (22), 4636–4656. 10.1021/jp402702u. [DOI] [PubMed] [Google Scholar]
- Mellouki A.; Wallington T. J.; Chen J. Atmospheric Chemistry of Oxygenated Volatile Organic Compounds: Impacts on Air Quality and Climate. Chem. Rev. 2015, 115 (10), 3984–4014. 10.1021/cr500549n. [DOI] [PubMed] [Google Scholar]
- Riedel T. P.; Wagner N. L.; Dubé W. P.; Middlebrook A. M.; Young C. J.; Öztürk F.; Bahreini R.; VandenBoer T. C.; Wolfe D. E.; Williams E. J.; et al. Chlorine Activation within Urban or Power Plant Plumes: Vertically Resolved ClNO2 and Cl2Measurements from a Tall Tower in a Polluted Continental Setting. J. Geophys. Res. 2013, 118 (15), 8702–8715. 10.1002/jgrd.50637. [DOI] [Google Scholar]
- Mellouki A.; Ammann M.; Cox R. A.; Crowley J. N.; Herrmann H.; Jenkin M. E.; McNeill V. F.; Troe J.; Wallington T. J. Evaluated Kinetic and Photochemical Data for Atmospheric Chemistry: Volume VIII – Gas-Phase Reactions of Organic Species with Four, or More, Carbon Atoms (≥C4). Atmos. Chem. Phys. 2021, 21 (6), 4797–4808. 10.5194/acp-21-4797-2021. [DOI] [Google Scholar]
- Mellouki A.; Oussar F.; Lun X.; Chakir A. Kinetics of the Reactions of the OH Radical with 2-Methyl-1-Propanol, 3-Methyl-1-Butanol and 3-Methyl-2-Butanol between 241 and 373 K. Phys. Chem. Chem. Phys. 2004, 6 (11), 2951–2955. 10.1039/b316514k. [DOI] [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]
- Nishino N.; Arey J.; Atkinson R. Rate Constants for the Gas-Phase Reactions of OH Radicals with a Series of C 6 – C 14 Alkenes at 299 ± 2 K. J. Phys. Chem. A 2009, 113 (5), 852–857. 10.1021/jp809305w. [DOI] [PubMed] [Google Scholar]
- Barbosa T. S.; Peirone S.; Barrera J. A.; Abrate J. P. A.; Lane S. I.; Arbilla G.; Bauerfeldt G. F. Rate Coefficients for the Reaction of OH Radicals with Cis-3-Hexene: An Experimental and Theoretical Study. Phys. Chem. Chem. Phys. 2015, 17 (14), 8714–8722. 10.1039/C4CP05760K. [DOI] [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]
- 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]
- 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]
- 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]
- 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]
- 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]
- 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. 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.