Previous Article in Journal
Studies on the Radziszewski Reaction—Synthesis and Characterization of New Imidazole Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydroxyl Radical-Initiated Reaction of Nerol: A Pathway to Secondary Pollutants in an Indoor Environment

by
Angappan Mano Priya
* and
Gisèle El Dib
*
IPR (Institut de Physique de Rennes)—UMR 6251, Université de Rennes, CNRS, F-35000 Rennes, France
*
Authors to whom correspondence should be addressed.
Reactions 2025, 6(3), 49; https://doi.org/10.3390/reactions6030049
Submission received: 7 July 2025 / Revised: 1 September 2025 / Accepted: 9 September 2025 / Published: 12 September 2025

Abstract

Nerol ((Z)-3,7-dimethylocta-2,6-dien-1-ol), (C10H18O), is a monoterpene alcohol that belongs to the family of BVOCs emitted naturally by means of vegetation and is found in various medicinal plants. This species attracted attention in the field of atmospheric chemistry due to its unique structural, chemical and environmental properties. In this work, OH-addition and H-abstraction reactions of Nerol by OH radical have been investigated using M06-2X, CBS-QB3 and CCSD(T) with 6-311++G(d,p) basis set. The OH addition at the C=C double bond of Nerol was shown to be the most favorable, with a small relative energy barrier of −6.86 kcal/mol and H-abstraction at the CH2 group exhibits a relative energy barrier of 0.08 kcal/mol at CCSD(T)/6-311++G(d,p) level of theory. The obtained overall rate coefficient at 298 K is 9.68 × 10−10 cm3 molecule−1 s−1 using canonical variational transition state theory with small curvature tunnelling method (CVT/SCT), which is in good agreement with the experimental rate coefficient determined by Mahecha et al. (kOH = (1.60 ± 0.2) × 10−10) at 296 ± 2 K. The obtained rate coefficient exhibits negative temperature dependence, and the atmospheric lifetime of Nerol is about 18 min. The predicted oxidation pathways reveal the formation of key products such as formaldehyde, glycolaldehyde and 6-Methyl-hept-5-en-2-ol, which is also observed in previous experimental studies, indicating good agreement between theoretical and experimental findings. This study constitutes the first theoretical study and its dependence on temperature exploration, offering detailed insights into the degradation pathways and environmental impact of Nerol initiated by hydroxyl radicals.

Graphical Abstract

1. Introduction

Biogenic volatile organic compounds (BVOCs) emitted by plants are the key components in the atmosphere and play a significant role in the formation of secondary organic aerosols (SOAs) [1]. BVOCs have significant effects not only within vegetation and organisms but also in the atmosphere at different spatiotemporal scales, mainly during heat waves and in densely populated areas [2,3]. Among BVOCs, the emission of terpenes and their Terpene Oxidation Products (TOPs) in the indoor air environment impacts atmospheric chemistry. Volatile biogenic terpenes in the indoor environment are crucial, and several studies report the indoor OH radical concentration order [4,5,6,7] as 105 molecules cm−3, whereas the outdoor OH radical concentration is typically about 106 molecules cm−3 in forested areas, as reported in earlier field studies [8,9]. The major atmospheric pollutants in the indoor environment are primarily emitted by humans, who act as a potent and mobile source of various chemicals. Several studies have shown that these TOPs are found with high levels in closed environments, particularly from cleaning products, air fresheners, perfumes and essential oils [10,11,12]. Once emitted into the atmosphere, terpenes are mainly removed by reacting with main atmospheric oxidants (OH, NO3, O3, Cl), leading to the formation of secondary species that are sometimes more toxic and harmful, including formaldehyde, ozone, carbon dioxide and SOAs, whose effects on human health and climate are well-known. The atmospheric degradation of terpenes has important implications in both indoor and outdoor environments. These TOPs reactions influence indoor air quality, and indoor SOA contributes significantly to the total human exposure burden as people spend ~90% of their time indoors [13,14,15].
Nerol ((Z)-3,7-dimethylocta-2,6-dien-1-ol) ((C10H18O)) is a monoterpene alcohol that belongs to the family of BVOC, which is emitted naturally by means of vegetation and found in various medicinal plants like Lippia spp. and Melissa officinalis L. [16,17,18]. Nerol is commonly used in essential oils and fragrance, flavour in soaps, detergents, shampoos, perfumes and in household cleaning products [19,20]. Nerol has widespread applications in cosmetics, pharmaceutical and medicinal uses, such as anti-inflammatory, anti-microbial and antioxidant, and it has potential anti-cancer effects [21,22]. Although Nerol has numerous applications, its degradation mechanism with OH radicals in the atmosphere are of significant interest due to its impact on air quality, which remains insufficiently explored, to the best of our knowledge. Due to its limited attention, herein, in this study, the first theoretical insight into the mechanistic and kinetic exploration of Nerol with the OH radical has been explored. Figure 1 depicts the chemical structure of Nerol. Nerol is a monoterpene alcohol that consists of two double bonds at C8=C4 and C5=C6, one hydroxyl group at C11 and methyl groups at C4 and C6. Nerol is presumed to undergo both H-abstraction and OH addition reactions with atmospheric OH radicals. In addition to these initial reaction pathways, it can subsequently react with O2, NO and HO2 radicals, leading to the formation of peroxy and alkoxy radical intermediates. To the best of our knowledge, there is no theoretical study of the reaction mechanism of Nerol with the OH radical. With respect to the available data on the kinetics of the reaction of Nerol with OH radical, experimental studies indicated that the reaction proceeds rapidly, with the rate coefficient typically in the range of 10−11 to 10−10 cm3 molecule−1 s−1 at 298 K under atmospheric conditions [18,23]. Using relative rate technique, G.L. Mahecha et al. [18] determined the rate coefficient of Nerol with OH radical 1.6 ± 0.2 × 10−10 cm3 molecule−1 s−1 at 298 K and there is only one experimental study on gas phase kinetics of Nerol with O3 by Alossaily et al. [23] over the temperature range of 298–353 K at atmospheric pressure using absolute and relative rate method by rigid atmospheric simulation chamber coupled to a proton transfer reaction mass spectrometer (PTR-ToF-MS). They found the rate coefficient of Nerol with ozone to be 8.89 ± 0.90 × 10−16 cm3 molecule−1 s−1 at 298 K. In addition to atmospheric oxidants, no study has been reported on Nerol with Cl and NO3 radicals.
Hence, this study investigates the atmospheric degradation of Nerol by OH radicals, emphasizing the significance of its transformation products and its estimated lifetime under typical atmospheric OH radical concentrations. Theoretical analysis is employed to assess the extent to which the molecular structure of Nerol facilitates its degradation, with particular focus on its potential role in secondary organic aerosols (SOA) formation and implications with respect to the function of temperature. Hence, a deeper understanding of its reaction pathways is essential for evaluating its role in secondary pollutant formation and for improving models of indoor and atmospheric chemistry.

2. Computational Details

Quantum chemical computations were performed to predict the reaction pathways. The minimum energy structures and transition states were located on the potential energy surface (PES) at the M06-2X/6-311++G(d,p) level of theory [24,25,26]. To ensure that each transition state smoothly connects the correct reactant and product minima, intrinsic reaction coordinate analysis (IRC) [27,28] calculations were carried out at the same level of theory. All the reactive species exhibited only real (positive) vibrational frequencies, confirming their nature as minima on the PES, except for the transition states, which exhibited a single imaginary frequency. To obtain more accurate energies, single-point calculations were performed at the CCSD(T) [29,30] level in conjunction with the 6-311++G(d,p) level of theory and the CBS-QB3 [31] level of theory. M06-2X is more efficient and reliable for TS geometries and vibrational analyses. CBS-QB3 gives thermochemical accuracy, whereas CCSD(T) is used for benchmark, single-point energetics for key stationary points. All the electronic structure calculations were performed using Gaussian 16 [32] and MOLPRO 2015 [33] program packages.
The rate coefficient of all the initial H-abstraction and OH-addition reaction of Nerol with OH radical are calculated using canonical variational transition state theory (CVT) with small curvature tunnelling method (CVT/SCT) with the temperature range of 278 to 350 K. Temperature (T) is given by the following relation from canonical variational transition state theory (CVT) [34,35,36]:
k C V T ( T ) = min s k G T ( T , s )
where
k G T ( T , s ) = σ k B T h Q G T ( T , s ) φ R ( T ) e v M E P ( s ) k B T
where k G T ( T , s ) is the Generalized Transition State (GTS) theory rate coefficient dividing the surface s, σ is the symmetry factor to illustrate the possibility of more than one symmetry-related reaction path, k B is Boltzmann’s coefficient, h is Planck’s coefficient, φ R ( T ) is the reactant classical partition function per unit volume, Q G T ( T , s ) is the classical partition function of a GTS and v M E P ( s ) is the classical potential energy at point s on the minimum energy path. The kinetics of all reaction pathways were calculated using the GAUSSRATE 2017B program [37], which is an interface between the GAUSSIAN 16 [32] and POLYRATE 2017 [38] programs.

3. Results and Discussion

3.1. Abstraction and Addition Pathways of Nerol with OH Radical

Nerol undergoes both H-abstraction and OH-addition mechanisms with the OH radical due to the presence of unsaturated alcohol (C=C) double bonds. The reaction scheme of H-abstraction and OH radical addition sites is shown in Scheme 1. The structures of all the species involved in H-abstraction and OH addition of Nerol with OH radical optimized at M06-2X/6-311++G(d,p) are given in the Supporting Information Figure S1. The unsaturated double bonds at C8=C4 at the terminal carbon (α-site) and C5=C6 of Nerol allow an OH-addition mechanism. Remaining OH, C-H, CH2 and CH3 groups of Nerol undergo H-abstraction reaction with OH radical. The relative energy profile of H-abstraction and OH-addition mechanism of Nerol with OH radical calculated at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p) is shown in Figure 2. Single-point energy is calculated at the CBS-QB3 and CCSD(T)/6-311++G(d,p) level of theory and the thermodynamic parameters enthalpy (ΔH) and Gibbs free energy (ΔG) calculated at M06−2X/6−311++G(d,p) are tabulated in Table 1. Harmonic vibrational frequencies of reactant (R), pre-reactive complex (RC), transition states (TS), intermediate complexes (IC), intermediates (I) and products (P) of Nerol with OH radical at the M06-2X/6-311++G(d,p) level of theory for H-abstraction and OH-addition pathways are given in Supplementary Information Table S1.
At the entrance channel, the reaction of Nerol with OH radical reactant (R) proceeds to form the pre-reactive complex (RC) with an intermolecular hydrogen bond O30-H31•••O1 in the form of barrierless reaction with a relative energy of −8.31 kcal/mol at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p). The H-abstraction and OH-addition mechanism of Nerol with OH radical undergoes the following steps for the formation of intermediates (I1 to I9) and the products (P1 to P4). (i) The reactant (R) leads to the formation of pre-reactive complex (RC), (ii) RC undergoes to form transition state (TS1–TS13), leading to the formation of intermediate complex of Nerol along with water molecule and products, and (iii) further intermediate complexes are decomposed into intermediates and H2O as shown in Figure S1. The H-abstraction pathways of Nerol with OH radical via OH, CH, CH2 and CH3 groups undergo transition state (TS1–TS9) from the pre-reactive complex (RC) with relative energy of 0.08–5.52 kcal/mol at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p). The obtained relative energy values are comparable with the CBS-QB3 level of theory depicted in Table 1. Similarly, at the CBS-QB3 level of theory, all nine H-atom abstraction pathways exhibit the relative energy of −0.25 to 2.22 kcal/mol for TS1–TS9. From the overall abstraction pathways of Nerol with OH radical, we found the H-abstraction from allylic -CH2 group at terminal carbon α-site is more dominant than all other H-abstraction pathways with small relative energy barrier of TS2 0.08 kcal/mol CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p), whereas in CBS-QB3 it shows a relative energy barrier of −2.59 kcal/mol. Even though the OH radical abstracts the H-atom of Nerol from different positions -OH, -CH, -CH2 and -CH3 groups through transition states TS1–TS9 from different C-atoms, TS2 is found to be more favorable with a small barrier. This kind of stable formation may be due to the presence of hydrogen bonding at the terminal carbon. From the optimized structures of Figure S1, it is well known that H31-O30•••H27 in TS2 has a hydrogen bond length of 1.58 Å, whereas the C11-H27 bond breaks with a bond length of 1.14 Å, leading to the formation of intermediate complex IC2 along with a water molecule. Further, the intermediate complex breaks into intermediate I2 and H2O. Remaining H-abstraction pathways from -OH, -CH2, -CH and -CH3 groups initiated via RC undergo transition state TS1, TS3–TS9, leading to the formation of intermediate complexes along with a water molecule. All the abstraction reaction mechanisms are exothermic and spontaneous in nature.
As we discussed earlier, H-abstraction of Nerol by the OH radical dominates at the allylic CH2 at the terminal carbon α-site. In the same manner, the reactive π bonds of Nerol undergo electrophilic addition of hydroxyl radical OH at its two double bonds, C6=C5 and C4=C8. OH radical addition at two double-bond positions of Nerol takes from the reactive complex RC through a transition state (TS10–TS13) with a relative energy barrier ranging from −4.35 to −6.86 kcal/mol at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p). As shown in Figure 2 and Scheme 1, addition leads to the formation of products P1 to P4. Among these four addition pathways C4=C8, the addition of OH radical at C4 is the most dominant pathway with a relative energy barrier of −6.86 kcal/mol via transition state TS11 leading to the formation of 3,7-Dimethyl-oct-6-ene-1,3-diol (P2) at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p), whereas at CBS-QB3, it exhibits −6.92 kcal/mol. The addition of the OH radical at the C4 position favors the predominant pathway due to less steric hindrance at this terminal carbon α-site. These kind of products acts as a precursor in atmospheric chemistry. From the PES Figure 2, the reported relative energy barrier for transition states TS10, TS12 and TS13 is less dominant than that of TS11. All four addition reaction pathways are exothermic and spontaneous in nature with respect to the thermodynamic parameters tabulated in Table 1. Hence, from the above discussion, we found OH-addition pathways are more dominant due to the presence of double bonds than H-abstraction pathways.

3.2. Competing Pathways of RO2 and Their Reactions with NO and HO2

The intermediates (I1 to I9) obtained via H-abstraction pathways of Nerol may undergo reaction with O2 to form a peroxy radical intermediate [39], which later may react with NO and HO2 to form possible intermediates and products as depicted in Scheme 2. The PES of peroxy radical intermediates and products are given in Figure 3, calculated at the M06−2X/6−311++G(d,p) level of theory, and the values of H-abstraction secondary reaction pathways are depicted in Table S2. The optimized structures of secondary abstraction pathways with O2, NO and HO2 are given in Figure S2. Initially, all the intermediates (I1 to I9) obtained from H-abstraction pathways undergo reaction with O2 to form a peroxy radical. The formation of peroxy radical RO2 and its interaction with NO proceed in the form of a barrierless reaction. These reactions are highly exothermic in nature. The reactions of NO with peroxy radicals are mainly responsible for atmospheric reactions, especially in polluted air [40,41]. From the Scheme 2, we found that the intermediate I1, which reacts with O2, leads to an intermediate I10. The intermediate I10, which reacts with NO, leads to the formation of I11 3,7-Dimethyl-octa-2,6-dien-1-yl-peroxide radical ( R O O + NO 2 ) through a transition state TS15 with a relative energy barrier of 42.46 kcal/mol. The reaction is endothermic and non-spontaneous with reaction enthalpy ΔH and Gibbs free energy ΔG as 17.70 and 17.88 kcal/mol, respectively, at the M06-2X level of theory (Table S2). Hence, from the PES Figure 3, we found that all the VOC intermediates reaction products with NO exhibit exothermic reaction in nature except intermediate I10, which reveals an endothermic nature with the release of NO2.
According to studies from the literature, we know VOCs undergo more competing pathways with other atmospheric oxidants [40,42,43] like NO and HO2. A schematic representation of the peroxy radical intermediate with HO2 is clearly depicted in Scheme 2. HO2 is an important intermediate species, and it plays a crucial role in oxidation reactions and radical propagation [44]. Hence, in this study, the peroxy intermediates interact with HO2 through H-atom abstraction by O2. Each individual peroxy intermediate interacts with HO2 via transition state (TS14–TS23) with a relative energy barrier varying from −0.74 to 24 kcal/mol, respectively, at the M06-2X level of theory, leading to product (P5 to P13), as shown in PES Figure 3 and to the formation of hydroperoxide (ROOH) product, along with O2. From the thermodynamic parameters, we found that all these reactions are exothermic and spontaneous in nature, and the values are tabulated in Table S2.
This study aims to offer a comprehensive insight into the full range of reaction pathways and mechanisms involved in Nerol with the OH radical. This is the first theoretical study to explore the entire reaction pathways of products (P1 to P4) of Nerol by OH-addition pathways, as shown in Scheme 3, and the thermodynamic parameters calculated at M06-2X/6-311++g(d,p) level of theory. The PES of all the secondary reactions of addition pathways is depicted in Figure 4. These products have been obtained through the addition of the OH radical to Nerol at the double-bond position. As a subsequent reaction, these products (P1 to P4) react with O2 in the form of a barrierless reaction. Product P1 reacts with O2 to lead to an intermediate (I28) 3-Hydroperoxy-3-methyl-oct-6-ene-1,2-diol radical with relative enthalpy and Gibbs free energy as −34.17 and −28.31 kcal/mol, respectively (Table 2). The addition of O2 takes place at the terminal carbon C4 as shown in Figure S2. In the same manner, the remaining three products, P2 to P4, lead to an intermediate 2-Hydroperoxy-3,7-dimethyl-oct-6-ene-1,3-diol radical (I30), 7-Hydroperoxy-3,7-dimethyl-oct-6-ene-1,6-diol radical (I32) and 6-Hydroperoxy-3,7-dimethyl-octane-1,7-diol radical (I34), respectively (Scheme 3). The PES of the obtained intermediates is given in Figure 4. All these reactions are exothermic in nature.
NO plays a significant role in radical chemistry, mainly in atmospheric reactions. When a peroxy radical intermediate reacts with NO, it undergoes rapid decomposition of R O , along with NO2. These kinds of intermediates are more unstable and play a crucial role in ozone formation in polluted environments [40]. As per our reaction, Scheme 3, all four peroxy intermediates undergo reaction with NO through a barrierless reaction, as shown in PES Figure 4. These reactions are highly exothermic and spontaneous in nature with high reaction enthalpy and Gibbs free energy as tabulated in Table 2. As we know from the above discussion, P2 is the most favorable reaction site in which the addition of OH occurs at the double bond of Nerol with the terminal carbon atom C4. Hence, P2 reacts with O2 in the form of a barrierless reaction leading to an intermediate I30 with reaction enthalpy and Gibbs free energy as −29.17 and −24.39 kcal/mol, respectively, at M06-2X/6-311++g(d,p) level of theory. Further, the intermediate I30 reacts with NO, leading to an intermediate (I31) 3,7-Dimethyl-oct-6-ene-1,2,3-triol radical with the release of NO2. This reaction is more unstable and occurs quickly due to high exothermic values of enthalpy and Gibbs free energy of −6.68 and −64.56 kcal/mol, respectively, at the M06-2X/6-311++g(d,p) level of theory. The obtained results are consistent with a previous experimental study conducted by Mahecha et al. [18], published in 2025. They obtained the reaction products in the presence of NOx, identified by solid-phase microextraction coupled to gas chromatography–mass spectrometry (SPME-GC–MS). They reported the yields of formaldehyde and acetone, which were found to be (0.06 ± 0.01) and (0.5 ± 0.1), respectively.

3.3. Mechanistic Study: Formation of Formaldehyde, Glycoaldehyde and 6-Methyl-hept-5-en-2-ol from Intermediate I31

It is worth noting and interesting to study the end products of the reaction mechanism of Nerol with OH radical in the presence of atmospheric oxidants like O2, NO and HO2 to know their atmospheric implications at the tropospheric level. To the best of our knowledge, there is no theoretical study on the investigation of Nerol oxidation products. Therefore, in order to investigate the entire mechanistic study for the most favorable reaction site of Nerol, we made an attempt to study the full decomposition mechanism. Scheme 4 shows the possible reaction pathways of intermediate I31, which is the most important intermediate compound expected under polluted conditions, with O2 and their thermodynamic values calculated at M06-2X/6-311++g(d,p) level of theory are tabulated in Table 2. In order to assess the atmospheric implications of Nerol, we made an attempt to explore the decomposition pathways of most of the favorable reactions. As an initial step of decomposition reaction, I31 undergoes cleavage of P14 (6-Methyl-hept-5-en-2-ol) and I36 (CH2OH-CHO) glycoaldehyde through a transition state TS24 with a relative energy barrier of 3.47 kcal/mol at M06-2X/6-311++g(d,p) level of theory. In transition state TS24, the C8=C4 double bond undergoes cleavage with the bond length of 2.08 Å to give their respective product (P14) and the intermediate (I36). This reaction is exothermic and spontaneous in nature with reaction enthalpy and Gibbs free energy of −0.03 and −2.78 kcal/mol, respectively, as shown in PES Figure 4. The formation of 6-methyl-hept-5-en-2-ol (P14) clearly supports the findings of a previous experimental study on the reaction of Nerol with OH radicals, as reported by Mahecha et al. [18]. Further, CH2OH-CHO glycoaldehyde (I36) undergoes reaction with O2 through a transition state TS25 in which the C2-H4 bond elongates with the bond length of 1.59 Å and the H4 atom is abstracted with O9 of the oxygen atom O2. The reaction occurs with a relative energy barrier of 26.92 kcal/mol through a transition state, TS25, at M06-2X/6-311++g(d,p) level of theory, leading to P15 (CH2OH-C=O) along with HO2.
As a next step, intermediate I31 undergoes bond cleavage of C11-C8 to give an intermediate I37 (2,6-Dimethyl-hept-5-ene-1,2-diol radical) and I38 (CH2OH). This reaction occurs through a transition state, TS26, in which C11-C8 bond cleavage takes place with the bond length of 2.09 Å with a relative energy barrier of 14.60 kcal/mol, as shown in PES Figure 4. Finally, the last major reaction pathway of intermediate I38 reacts with O2 present in the atmosphere, in which the H atom is abstracted with O2, leading to stable product P16 formaldehyde (HCHO) along with HO2. The formation of formaldehyde occurs through a transition state TS27, in which the O1-H5 bond elongates with a bond length of 1.03 Å and the H5 atom is abstracted with O6 with a bond length of 1.43 Å, with a relative energy barrier of −3.57 kcal/mol. The obtained product is highly exothermic and spontaneous in nature with reaction enthalpy and Gibbs free energy as −38.05, −34.47 kcal/mol, respectively, at M06-2X/6-311++g(d,p) level of theory. The predicted oxidation pathway of formaldehyde acts as a key product, indicating a good agreement between theoretical and experimental findings. Among the possible entrance channel pathways, OH addition to the C=C double bond of Nerol is by far the dominant process, more than H-abstraction pathways.

3.4. Rate Coefficient, Branching Ratio of Nerol with OH Radical

The rate coefficients were determined theoretically using canonical variational transition state theory with small curvature tunnelling method (CVT/SCT) with the temperature range of 278–350 K, 1 atmospheric pressure, using M06-2X/6-311++g(d,p) level of theory, and their values are tabulated in Table 3. Based on the obtained rate values, we found that OH addition is more favorable than H-abstraction. Tunnelling factors are included for both H-abstraction and OH-addition pathways. The formation of product P2 is more kinetically favorable with a rate coefficient of 6.34 × 10−10 cm3 molecule−1 s−1 at 298 K. Hence, the overall rate coefficient of Nerol with OH radical for the most favorable pathway P2 exhibits 9.68 × 10−10 cm3 molecule−1 s−1. The rate coefficient shows negative temperature dependence; as the temperature increases, the rate coefficient decreases. This might be due to the formation of pre-reactive complexes in radical reactions and is well agreed with our previous theoretical studies and the literature [11,45]. Similarly, the reactions of α-pinene, myrcene and limonene with OH radicals also exhibit strong negative temperature dependence, consistent with OH addition dominating their entrance channels [13,14,15]. The three-parameter Arrhenius expression for all the pathways is given in the Supplementary Information, Table S3. As shown in Table 3, the rate coefficients obtained for H-abstraction pathways exhibit a (10−11 to 10−14) order of magnitude for the whole temperature range of 278–350 K. The obtained rate coefficient results of Nerol with OH are comparable with previous experimental study determined by Mahecha et al. [18], where a rate coefficient of 1.60 ± 0.2 × 10−10 cm3 molecule−1 s−1 was obtained at 298 K. Similarly, among the H-abstraction pathways, we found pathway 2, H-abstraction at terminal carbon is the most dominant pathway with small energy barrier, which is thermodynamically and kinetically favorable with 3.88 × 10−11 cm3 molecule−1 s−1 at 298 K. The overall three-parameter Arrhenius expression is given below in Equation (3).
k = 4.14 × 10−13 Exp (2308.6/T)
The contribution of each reaction channel is calculated theoretically using Equation (4) given below:
Branching   ratio   =   k k t o t a l × 100
Table 4 depicts branching ratios of H-atom abstraction and OH addition reaction pathways of Nerol with the OH at M06-2X/6-311++G(d,p) level of theory over the temperature range of 278–350 K. The branching ratios of all reaction channels were evaluated, revealing that Pathway 11 is the dominant contributor, with a major contribution of 65.48% at 298 K. The addition pathways at the double bonds of Nerol contribute more significantly than the whole H-atom abstraction pathways.

4. Evaluating the Role of Nerol in Atmospheric Chemistry and Conclusions

This study represents the first theoretical investigation into the OH-initiated atmospheric degradation of Nerol. Our computational results elucidate the primary reaction pathways, key intermediates and products contributing to the formation of secondary organic aerosol (SOA). The calculated energetics and rate coefficients indicate that the OH-addition mechanism is more kinetically and thermodynamically favorable than the H-abstraction pathways of Nerol under atmospheric conditions. These results give valuable insights into monoterpenoid degradation in atmospheric chemistry. As we know, terpenes are significant contributors to global VOCs emissions, suggesting that monoterpene alone accounts for about 10.9% of total global VOCs emissions [3]. The atmospheric lifetime of Nerol with OH radical is ~18 min at the local scale by using the OH radical concentration [46] of 1 × 106 molecule cm−3 using the relation given below in Equation (4).
τ O H = 1 k O H [ O H ]
Indeed, emissions of terpenes increase with global warming, thereby enhancing the formation of greenhouse gases (GHGs) and secondary organic aerosols (SOAs), which further contribute to climate change. However, regarding terpene emissions, the atmospheric lifetime of Nerol due to its reaction with OH radicals is found to be very short. This indicates that Nerol has a limited global impact, as it is rapidly removed from the atmosphere with an impact at the local scale.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/reactions6030049/s1, Figure S1. Optimized structures of H-abstraction and OH-addition of Nerol by OH radical calculated at M06-2X/6-311++g(d,p) level of theory. Figure S2. Optimized structures of intermediates with O2, NO and HO2 calculated at M06-2X/6-311++g(d,p) level of theory. Table S1. Harmonic vibrational frequencies of reactant (R), prereactive complexe (RC), transition states (TS), intermediate complexes (IC), intermediates (I) and products (P) of Nerol with OH radical at M06-2X/6-311++G(d,p) level of theory for H-abstraction and OH addition pathways. All the frequencies are in cm-1. Table S2. The relative energy, enthalpy and Gibb’s free energy (in kcal/mol) for the secondary reaction of intermediates with O2 and HO2 radical calculated at M06-2X/6-311++g(d,p) level of theory. Table S3. Three parameter Arrhenius expression for all the pathways.

Author Contributions

Conceptualization, A.M.P.; methodology, A.M.P.; software A.M.P.; validation, A.M.P. and G.E.D.; formal analysis, A.M.P. and G.E.D.; investigation, A.M.P. and G.E.D.; resources, A.M.P. and G.E.D.; data curation, A.M.P.; writing—original draft preparation, A.M.P. and G.E.D.; writing—review and editing, A.M.P. and G.E.D.; visualization, A.M.P. and G.E.D.; supervision, G.E.D.; project administration, G.E.D.; funding acquisition, G.E.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors gratefully thank the SAD-Brittany council program for granting project funding to accomplish this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hirose, S.; Satake, A. Theoretical analyses for the evolution of biogenic volatile organic compounds (BVOC) emission strategy. Ecol. Evol. 2024, 14, e11548. [Google Scholar] [CrossRef]
  2. Gu, S.; Guenther, A.; Faiola, C. Effects of Anthropogenic and Biogenic Volatile Organic Compounds on Los Angeles Air Quality. Environ. Sci. Technol. 2021, 55, 12191–12201. [Google Scholar] [CrossRef]
  3. Guenther, A.B.; Jiang, X.; Heald, C.L.; Sakulyanontvittaya, T.; Duhl, T.; Emmons, L.K.; Wang, X. The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1): An extended and updated framework for modeling biogenic emissions. Geosci. Model Dev. 2012, 5, 1471–1492. [Google Scholar] [CrossRef]
  4. Carslaw, N.; Fletcher, L.; Heard, D.; Ingham, T.; Walker, H. Significant OH production under surface cleaning and air cleaning conditions: Impact on indoor air quality. Indoor Air 2017, 27, 1091–1100. [Google Scholar] [CrossRef]
  5. Wang, N.; Zannoni, N.; Ernle, L.; Bekö, G.; Wargocki, P.; Li, M.; Weschler, C.J.; Williams, J. Total OH Reactivity of Emissions from Humans: In Situ Measurement and Budget Analysis. Environ. Sci. Technol. 2021, 55, 149–159. [Google Scholar] [CrossRef]
  6. Weschler, C.J.; Shields, H.C. Production of the Hydroxyl Radical in Indoor Air. Environ. Sci. Technol. 1996, 30, 3250–3258. [Google Scholar] [CrossRef]
  7. Weschler, C.J.; Shields, H.C. Measurements of the Hydroxyl Radical in a Manipulated but Realistic Indoor Environment. Environ. Sci. Technol. 1997, 31, 3719–3722. [Google Scholar] [CrossRef]
  8. Lelieveld, J.; Butler, T.M.; Crowley, J.N.; Dillon, T.J.; Fischer, H.; Ganzeveld, L.; Harder, H.; Lawrence, M.G.; Martinez, M.; Taraborrelli, D.; et al. Atmospheric oxidation capacity sustained by a tropical forest. Nature 2008, 452, 737–740. [Google Scholar] [CrossRef] [PubMed]
  9. Whalley, L.K.; Edwards, P.M.; Furneaux, K.L.; Goddard, A.; Ingham, T.; Evans, M.J.; Stone, D.; Hopkins, J.R.; Jones, C.E.; Karunaharan, A.; et al. Quantifying the magnitude of a missing hydroxyl radical source in a tropical rainforest. Atmos. Chem. Phys. 2011, 11, 7223–7233. [Google Scholar] [CrossRef]
  10. Calogirou, A.; Larsen, B.R.; Kotzias, D. Gas-phase terpene oxidation products: A review. Atmos. Environ. 1999, 33, 1423–1439. [Google Scholar] [CrossRef]
  11. Mano Priya, A.; El Dib, G. Mechanistic and kinetic study of limona ketone oxidation initiated by hydroxyl radical: Impact of indoor air pollution. New J. Chem. 2024, 48, 3036–3044. [Google Scholar] [CrossRef]
  12. Wolkoff, P.; Larsen, S.T.; Hammer, M.; Kofoed-Sørensen, V.; Clausen, P.A.; Nielsen, G.D. Human reference values for acute airway effects of five common ozone-initiated terpene reaction products in indoor air. Toxicol. Lett. 2013, 216, 54–64. [Google Scholar] [CrossRef]
  13. Braure, T.; Bedjanian, Y.; Romanias, M.N.; Morin, J.; Riffault, V.; Tomas, A.; Coddeville, P. Experimental Study of the Reactions of Limonene with OH and OD Radicals: Kinetics and Products. J. Phys. Chem. A 2014, 118, 9482–9490. [Google Scholar] [CrossRef]
  14. Chuong, B.; Davis, M.; Edwards, M.; Stevens, P.S. Measurements of the kinetics of the OH + α-pinene and OH + β-pinene reactions at low pressure. Int. J. Chem. Kinet. 2002, 34, 300–308. [Google Scholar] [CrossRef]
  15. Tan, Z.; Hantschke, L.; Kaminski, M.; Acir, I.H.; Bohn, B.; Cho, C.; Dorn, H.P.; Li, X.; Novelli, A.; Nehr, S.; et al. Atmospheric photo-oxidation of myrcene: OH reaction rate constant, gas-phase oxidation products and radical budgets. Atmos. Chem. Phys. 2021, 21, 16067–16091. [Google Scholar] [CrossRef]
  16. Durello, R.S.; Silva, L.M.; Bogusz, S. Química do Lúpulo. Química Nova 2019, 42, 900–919. [Google Scholar] [CrossRef]
  17. Escobar, P.; Milena Leal, S.; Herrera, L.V.; Martinez, J.R.; Stashenko, E. Chemical composition and antiprotozoal activities of Colombian Lippia spp. essential oils and their major components. Memórias Do Inst. Oswaldo Cruz 2010, 105, 184–190. [Google Scholar] [CrossRef]
  18. Mahecha, G.L.; Echeverria, G.C.; Barrera, J.A.; Garavagno, M.d.l.A.; Pino, G.A. Tropospheric degradation of nerol: Kinetics with OH radicals, product distribution in the presence of NOx and atmospheric implications. Atmos. Environ. 2025, 346, 121055. [Google Scholar] [CrossRef]
  19. Coêlho, M.L.; Islam, M.T.; Oliveira, G.L.d.S.; de Alencar, M.V.O.B.; Santos, J.V.d.O.; dos Reis, A.C.; da Mata, A.M.O.F.; Correia Jardim Paz, M.F.; Docea, A.O.; Calina, D.; et al. Cytotoxic and Antioxidant Properties of Natural Bioactive Monoterpenes Nerol, Estragole, and 3,7-Dimethyl-1-Octanol. Adv. Pharmacol. Pharm. Sci. 2022, 2022, 8002766. [Google Scholar] [CrossRef] [PubMed]
  20. Lapczynski, A.; Foxenberg, R.J.; Bhatia, S.P.; Letizia, C.S.; Api, A.M. Fragrance material review on nerol. Food Chem. Toxicol. 2008, 46, S241–S244. [Google Scholar] [CrossRef] [PubMed]
  21. Silva, G.d.S.e.; Marques, J.N.d.J.; Linhares, E.P.M.; Bonora, C.M.; Costa, É.T.; Saraiva, M.F. Review of anticancer activity of monoterpenoids: Geraniol, nerol, geranial and neral. Chem.-Biol. Interact. 2022, 362, 109994. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Z.; Yang, K.; Chen, L.; Yan, R.; Qu, S.; Li, Y.; Liu, M.; Zeng, H.; Tian, J. Activities of Nerol, a natural plant active ingredient, against Candida albicans in vitro and in vivo. Appl. Microbiol. Biotechnol. 2020, 104, 5039–5052. [Google Scholar] [CrossRef]
  23. Alossaily, M.G.; Shamas, M.; Chakir, A.; Roth, E. Temperature-Dependent Kinetic Study of the Gas Phase Ozonolysis of Linalool, Nerol, and Citronellol. Int. J. Chem. Kinet. 2025, 57, 342–350. [Google Scholar] [CrossRef]
  24. Aazaad, B.; Mano Priya, A.; Subramani, R. Environmental implications of oxalic and malonic acids with tropospheric oxidants. J. Phys. Org. Chem. 2024, 37, e4639. [Google Scholar] [CrossRef]
  25. Zhao, Y.; Truhlar, D.G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. [Google Scholar] [CrossRef]
  26. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  27. Gonzalez, C.; Schlegel, H.B. An improved algorithm for reaction path following. J. Chem. Phys. 1989, 90, 2154–2161. [Google Scholar] [CrossRef]
  28. Gonzalez, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  29. Noga, J.; Bartlett, R.J. The full CCSDT model for molecular electronic structure. J. Chem. Phys. 1987, 86, 7041–7050. [Google Scholar] [CrossRef]
  30. Ramabhadran, R.O.; Raghavachari, K. Extrapolation to the Gold-Standard in Quantum Chemistry: Computationally Efficient and Accurate CCSD(T) Energies for Large Molecules Using an Automated Thermochemical Hierarchy. J. Chem. Theory Comput. 2013, 9, 3986–3994. [Google Scholar] [CrossRef] [PubMed]
  31. Montgomery, J.A., Jr.; Frisch, M.J.; Ochterski, J.W.; Petersson, G.A. A complete basis set model chemistry. VI. Use of density functional geometries and frequencies. J. Chem. Phys. 1999, 110, 2822–2827. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. C.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  33. Werner, H.-J.; Knowles, P.J.; Celani, P.; Györffy, W.; Hesselmann, A.; Kats, D.; Knizia, G.; Köhn, A.; Korona, T.; Kreplin, D.; et al. MOLPRO, a Package of Ab Initio Programs, version 2015; University of Cardiff: Cardiff, UK, 2015. [Google Scholar]
  34. Garrett, B.C.; Truhlar, D.G. Generalized transition state theory. Bond energy-bond order method for canonical variational calculations with application to hydrogen atom transfer reactions. J. Am. Chem. Soc. 1979, 101, 4534–4548. [Google Scholar] [CrossRef]
  35. Garrett, B.C.; Truhlar, D.G. Criterion of minimum state density in the transition state theory of bimolecular reactions. J. Chem. Phys. 1979, 70, 1593–1598. [Google Scholar] [CrossRef]
  36. Garrett, B.C.; Truhlar, D.G.; Grev, R.S.; Magnuson, A.W. Improved treatment of threshold contributions in variational transition-state theory. J. Phys. Chem. 1980, 84, 1730–1748. [Google Scholar] [CrossRef]
  37. Zheng, J.; Zhang, S.; Corchado, J.; Chuang, Y.; Coitino, E.; Ellingson, B.; Truhlar, D. Gaussrate, Version 2017-B; University of Minnesota: Minneapolis, MN, USA, 2010. [Google Scholar]
  38. Zheng, J.; Zhang, S.; Lynch, B.; Corchado, J.; Chuang, Y.; Fast, P.; Hu, W.; Liu, Y.; Lynch, G.; Nguyen, K. Polyrate, version 2017; University of Minnesota: Minneapolis, MN, USA, 2010. [Google Scholar]
  39. Ingold, K.U. Peroxy radicals. Acc. Chem. Res. 1969, 2, 1–9. [Google Scholar] [CrossRef]
  40. Lightfoot, P.D.; Cox, R.A.; Crowley, J.N.; Destriau, M.; Hayman, G.D.; Jenkin, M.E.; Moortgat, G.K.; Zabel, F. Organic peroxy radicals: Kinetics, spectroscopy and tropospheric chemistry. Atmos. Environ. 1992, 26, 1805–1961. [Google Scholar] [CrossRef]
  41. Zhang, J.; Dransfield, T.; Donahue, N.M. On the Mechanism for Nitrate Formation via the Peroxy Radical + NO Reaction. J. Phys. Chem. A 2004, 108, 9082–9095. [Google Scholar] [CrossRef]
  42. Atkinson, R. Atmospheric chemistry of VOCs and NOx. Atmos. Environ. 2000, 34, 2063–2101. [Google Scholar] [CrossRef]
  43. Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103, 4605–4638. [Google Scholar] [CrossRef] [PubMed]
  44. Monks, P.S. Gas-phase radical chemistry in the troposphere. Chem. Soc. Rev. 2005, 34, 376–395. [Google Scholar] [CrossRef] [PubMed]
  45. Alvarez-Idaboy, J.R.; Mora-Diez, N.; Boyd, R.J.; Vivier-Bunge, A. On the Importance of Prereactive Complexes in Molecule-Radical Reactions:  Hydrogen Abstraction from Aldehydes by OH. J. Am. Chem. Soc. 2001, 123, 2018–2024. [Google Scholar] [CrossRef] [PubMed]
  46. Shu, Y.; Atkinson, R. Atmospheric lifetimes and fates of a series of sesquiterpenes. J. Geophys. Res. Atmos. 1995, 100, 7275–7281. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of Nerol.
Figure 1. Chemical structure of Nerol.
Reactions 06 00049 g001
Scheme 1. Possible reaction pathways of H-atom abstraction and OH radical addition reaction to the two C=C double bonds of Nerol with OH radical.
Scheme 1. Possible reaction pathways of H-atom abstraction and OH radical addition reaction to the two C=C double bonds of Nerol with OH radical.
Reactions 06 00049 sch001
Figure 2. Relative energy profile for the reaction of Nerol with OH radical via H-abstraction and OH addition at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p).
Figure 2. Relative energy profile for the reaction of Nerol with OH radical via H-abstraction and OH addition at CCSD(T)/6−311++G(d,p)//M06−2X/6−311++G(d,p).
Reactions 06 00049 g002
Scheme 2. Possible reaction pathways of H-atom abstracted intermediates (I1 to I9) reacting with O2, HO2 and NO to form peroxy radicals and stable products.
Scheme 2. Possible reaction pathways of H-atom abstracted intermediates (I1 to I9) reacting with O2, HO2 and NO to form peroxy radicals and stable products.
Reactions 06 00049 sch002aReactions 06 00049 sch002bReactions 06 00049 sch002c
Figure 3. Relative energy profile for the secondary reaction of OH-abstraction pathways with O2, NO and HO2 calculated at M06−2X/6−311++G(d,p).
Figure 3. Relative energy profile for the secondary reaction of OH-abstraction pathways with O2, NO and HO2 calculated at M06−2X/6−311++G(d,p).
Reactions 06 00049 g003
Scheme 3. Possible reaction pathways of OH radical addition reaction sites (P1 to P4) with O2 and NO to form peroxy radicals and intermediates.
Scheme 3. Possible reaction pathways of OH radical addition reaction sites (P1 to P4) with O2 and NO to form peroxy radicals and intermediates.
Reactions 06 00049 sch003
Figure 4. Relative energy profile for the secondary reaction of OH-addition pathways with O2, NO and their decomposition calculated at M06−2X/6−311++G(d,p).
Figure 4. Relative energy profile for the secondary reaction of OH-addition pathways with O2, NO and their decomposition calculated at M06−2X/6−311++G(d,p).
Reactions 06 00049 g004
Scheme 4. Possible reaction pathways of intermediate I31 with O2 to form formaldehyde (HCHO).
Scheme 4. Possible reaction pathways of intermediate I31 with O2 to form formaldehyde (HCHO).
Reactions 06 00049 sch004
Table 1. The relative energy, enthalpy and Gibbs free energy (in kcal/mol) for the reaction of Nerol with OH radical (H-abstraction and OH-addition) pathways calculated at M06-2X/6-311++G(d,p), CCSD(T) and CBS-QB3 level of theory.
Table 1. The relative energy, enthalpy and Gibbs free energy (in kcal/mol) for the reaction of Nerol with OH radical (H-abstraction and OH-addition) pathways calculated at M06-2X/6-311++G(d,p), CCSD(T) and CBS-QB3 level of theory.
REACTION PATHWAYSM06-2XCCSD(T)CBS-QB3
ΔEΔHΔGΔEΔE
PATHWAY 1R00000
RC1−10.81−8.84−0.34−8.31−6.32
TS11.130.118.543.832.22
IC1−19.71−18.96−12.38−17.22−16.95
I1+H2O−13.40−14.27−15.03−11.79−14.13
PATHWAY 2TS2−2.09−2.735.930.08−2.59
IC2−45.52−44.43−38.00−41.75−44.01
I2+H2O−37.14−37.87−38.59−33.78−39.57
PATHWAY 3TS31.21−0.368.593.43−0.93
IC3−16.19−14.51−8.32−12.83−12.92
I3+H2O−8.17−8.44−10.24−6.49−9.42
PATHWAY 4TS40.90−0.158.052.35−0.90
IC4−37.16−35.68−27.51−34.88−36.25
I4+H2O−28.66−29.27−29.88−27.16−31.98
PATHWAY 5TS5−0.66−1.568.421.15−2.38
IC5−31.72−32.09−28.16−29.02−35.98
I5+H2O−30.47−31.60−32.65−25.22−33.70
PATHWAY 6TS6−1.24−2.296.710.61−1.95
IC6−38.94−38.38−33.21−36.30−38.58
I6+H2O−32.97−33.77−35.35−30.82−36.06
PATHWAY 7TS71.950.538.843.69−0.28
IC7−16.54−15.46−9.79−13.48−13.55
I7+H2O−9.85−10.08−11.81−8.07−10.89
PATHWAY 8TS81.960.969.093.17−0.25
IC8−34.57−33.77−26.79−31.61−33.08
I8+H2O−28.63−29.14−30.33−27.30−31.73
PATHWAY 9TS93.622.1810.425.522.10
IC9−33.65−32.89−26.59−31.39−33.26
I9+H2O−28.76−29.16−29.68−27.08−31.54
PATHWAY 10TS10−6.61−4.914.83−4.45−5.93
P1−39.04−35.28−25.56−33.90−33.99
PATHWAY 11TS11−8.04−5.834.76−6.86−6.92
P2−39.00−35.28−24.22−34.26−33.61
PATHWAY 12TS12−5.17−3.546.13−2.81−3.37
P3−35.64−32.08−22.85−31.11−30.46
PATHWAY 13TS13−5.99−4.415.12−4.35−4.91
P4−34.80−31.74−21.54−30.80−31.96
Table 2. The relative energy, enthalpy and Gibbs free energy (in kcal/mol) for the reaction of secondary reactions of OH-addition pathways with O2, NO, calculated at M06-2X/6-311++G(d,p) level of theory.
Table 2. The relative energy, enthalpy and Gibbs free energy (in kcal/mol) for the reaction of secondary reactions of OH-addition pathways with O2, NO, calculated at M06-2X/6-311++G(d,p) level of theory.
Reaction PathwaysΔEΔHΔG
P1+O2000
I28−37.62−34.17−28.31
I28+NO000
I29−60.79−58.86−59.46
P2+O2000
I30−33.13−29.17−24.39
I30+NO000
I31−67.74−65.68−64.56
P3+O2000
I32−36.49−32.86−26.28
I32+NO000
I33−63.83−62.00−61.25
P4+O2000
I34−36.74−32.90−28.54
I34+NO000
I35−66.72−64.97−64.92
I31000
TS243.472.742.69
P14+I361.52−0.03−2.78
I36+O2000
TS2526.9225.6829.45
P15+HO28.576.78−1.68
I31000
TS2614.6013.1312.54
I37+I384.341.96−1.41
I38+O2000
TS27−3.57−5.75−2.18
P16+HO2−39.62−38.05−34.47
Table 3. Rate coefficient of H-atom abstraction and OH addition reaction pathways of Nerol calculated at CVT/SCT method at M06-2X/6-311++G(d,p) level of theory (cm3molecule−1s−1).
Table 3. Rate coefficient of H-atom abstraction and OH addition reaction pathways of Nerol calculated at CVT/SCT method at M06-2X/6-311++G(d,p) level of theory (cm3molecule−1s−1).
T (K)kI1
(10−13)
kI2kI3
(10−13)
kI4kI5
(10−12)
kI6
(10−12)
kI7
(10−13)
kI8
(10−14)
kI9
(10−14)
kII10kI11kI12
(10−11)
kI13Overall
2784.495.19 × 10−112.081.03 × 10−121.537.251.945.981.352.11 × 10−101.11 × 10−99.012.00 × 10−101.67 × 10−9
2884.644.46 × 10−112.151.07 × 10−121.426.732.096.361.511.47 × 10−108.31 × 10−107.371.54 × 10−101.26 × 10−9
2984.803.88 × 10−112.231.12 × 10−121.326.292.247.161.691.05 × 10−106.34 × 10−106.121.20 × 10−109.68 × 10−10
3084.93.41 × 10−112.301.16 × 10−121.245.921.467.541.887.63 × 10−114.93 × 10−102.529.53 × 10−117.33 × 10−10
3185.143.02 × 10−112.384.75 × 10−131.175.611.027.932.095.69 × 10−113.90 × 10−102.137.70 × 10−115.84 × 10−10
3285.312.70 × 10−112.464.80 × 10−131.115.331.098.322.314.32 × 10−113.14 × 10−101.836.31 × 10−114.74 × 10−10
3385.492.44 × 10−112.554.86 × 10−131.065.101.178.732.553.34 × 10−112.56 × 10−101.585.24 × 10−113.90 × 10−10
3483.489.75 × 10−122.644.92 × 10−131.024.901.259.142.812.63 × 10−112.12 × 10−101.384.41 × 10−113.13 × 10−10
3503.509.57 × 10−122.664.94 × 10−131.014.861.269.232.862.51 × 10−112.04 × 10−101.354.27 × 10−113.02 × 10−10
Table 4. Branching ratios of H-atom abstraction and OH addition reaction pathways of Nerol at M06-2X/6-311++G(d,p) level of theory.
Table 4. Branching ratios of H-atom abstraction and OH addition reaction pathways of Nerol at M06-2X/6-311++G(d,p) level of theory.
T (K)kI1kI2kI3kI4kI5kI6kI7kI8kI9kI10kI11kI12kI13Overall
2780.033.100.010.060.090.430.010.000.0012.6166.305.3811.97100.00
2880.043.540.020.090.110.530.020.010.0011.6465.975.8512.19100.00
2980.054.000.020.120.140.650.020.010.0010.7965.486.3312.39100.00
3080.074.650.030.160.170.810.020.010.0010.4067.253.4313.00100.00
3180.095.180.040.080.200.960.020.010.009.7466.833.6513.20100.00
3280.115.710.050.100.231.130.020.020.009.1366.313.8613.33100.00
3380.146.250.070.120.271.310.030.020.018.5865.684.0613.46100.00
3480.113.110.080.160.321.560.040.030.018.3967.684.4214.08100.00
3500.123.170.090.160.331.610.040.030.018.3167.544.4614.12100.00
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mano Priya, A.; El Dib, G. Hydroxyl Radical-Initiated Reaction of Nerol: A Pathway to Secondary Pollutants in an Indoor Environment. Reactions 2025, 6, 49. https://doi.org/10.3390/reactions6030049

AMA Style

Mano Priya A, El Dib G. Hydroxyl Radical-Initiated Reaction of Nerol: A Pathway to Secondary Pollutants in an Indoor Environment. Reactions. 2025; 6(3):49. https://doi.org/10.3390/reactions6030049

Chicago/Turabian Style

Mano Priya, Angappan, and Gisèle El Dib. 2025. "Hydroxyl Radical-Initiated Reaction of Nerol: A Pathway to Secondary Pollutants in an Indoor Environment" Reactions 6, no. 3: 49. https://doi.org/10.3390/reactions6030049

APA Style

Mano Priya, A., & El Dib, G. (2025). Hydroxyl Radical-Initiated Reaction of Nerol: A Pathway to Secondary Pollutants in an Indoor Environment. Reactions, 6(3), 49. https://doi.org/10.3390/reactions6030049

Article Metrics

Back to TopTop