Satya P.
Joshi
,
Timo T.
Pekkanen
,
Prasenjit
Seal
,
Raimo S.
Timonen
and
Arkke J.
Eskola
*
Molecular Science, Department of Chemistry, University of Helsinki, P.O. Box 55 (A.I. Virtasen aukio 1), FIN-00014, Helsinki, Finland. E-mail: arkke.eskola@helsinki.fi
First published on 2nd August 2021
The kinetics of the reaction between resonance-stabilized (CH3)2CCHCH2 radical (R) and O2 has been investigated using photoionization mass spectrometry, and master equation (ME) simulations were performed to support the experimental results. The kinetic measurements of the (CH3)2CCHCH2 + O2 reaction (1) were carried out at low helium bath-gas pressures (0.2–5.7 Torr) and over a wide temperature range (238–660 K). Under low temperature (238–298 K) conditions, the pressure-dependent bimolecular association reaction R + O2 → ROO determines kinetics, until at an intermediate temperature range (325–373 K) the ROO adduct becomes thermally unstable and increasingly dissociates back to the reactants with increasing temperature. The initial association of O2 with (CH3)2CCHCH2 radical occurs on two distinct sites: terminal 1(t) and non-terminal 1(nt) sites on R, leading to the barrierless formation of ROO(t) and ROO(nt) adducts, respectively. Important for autoignition modelling of olefinic compounds, bimolecular reaction channels appear to open for the R + O2 reaction at high temperatures (T > 500 K) and pressure-independent bimolecular rate coefficients of reaction (1) with a weak positive temperature dependence, (2.8–4.6) × 10−15 cm3 molecule−1 s−1, were measured in the temperature range of 500–660 K. At a temperature of 501 K, a product signal of reaction (1) was observed at m/z = 68, probably originating from isoprene. To explore the reaction mechanism of reaction (1), quantum chemical calculations and ME simulations were performed. According to the ME simulations, without any adjustment to energies, the most important and second most important product channels at the high temperatures are isoprene + HO2 (yield > 91%) and (2R/S)-3-methyl-1,2-epoxybut-3-ene + OH (yield < 8%). After modest adjustments to ROO(t) and ROO(nt) well-depths (∼0.7 kcal mol−1 each) and barrier height for the transition state associated with the kinetically most dominant channel, R + O2 → isoprene + HO2 (∼2.2 kcal mol−1), the ME model was able to reproduce the experimental findings. Modified Arrhenius expressions for the kinetically important reaction channels are enclosed to facilitate the use of current results in combustion models.
In this work, we focus on the oxidation kinetics of the allyl-type, resonance-stabilized (CH3)2CCHCH2 (2-methylbut-3-en-2-yl) radical. A potential source of this radical is a H-atom abstraction from 2-methyl-2-butene (2M2B) as well as pyrolysis and decomposition reactions of larger olefinic fuel constituents. 2M2B is a potential surrogate fuel for the combustion modeling of real fuels. As discussed by Westbrook et al.,16 under typical combustion conditions H-atom abstraction reactions of small radical species (e.g. H, O, OH, CH3, CH3OO etc.) from 2M2B is an important, if not crucial, route to initiate 2M2B oxidation. These reactions abstract H-atoms by breaking weak allylic C–H bonds (BDE ∼89 kcal mol−1) much more likely than the stronger vinylic C–H bond (BDE ∼108 kcal mol−1), resulting in production of two distinct allyl-type pentenyl radicals, as shown in Scheme 1.
![]() | ||
Scheme 1 Two allyl-type pentenyl radicals, CH2(CH3)CCHCH3 and (CH3)2CCHCH2, produced via H-atom abstraction reactions from 2-methyl-2-butene (2M2B). |
Both these allyl-type radicals, CH2(CH3)CCHCH3 and (CH3)2CCHCH2, are resonantly stabilized. Their (and allyl-type radical RA in general) reaction with molecular oxygen (O2) has a characteristically shallow RA + O2 → RAOO barrierless association reaction well-depth (∼20 kcal mol−119,24 compared to ∼40 kcal mol−1 for vinyl-type radicals25,26). This enables thermal decomposition of the RAOO adduct back to the reactants to start to compete with the forward association reaction already at relatively low temperatures, establishing the chemical equilibrium RA + O2 ⇌ RAOO. Upon increasing temperature, this may eventually lead to conditions with almost no net reaction, until a new reaction channel for the RA + O2 reaction, potentially, but not necessarily, becomes accessible at higher temperatures. For example, even at temperatures as high as 700 K, only an upper limit of 2 × 10−16 cm3 molecule−1 s−1 for the bimolecular rate coefficient of the CH3CHCHCH2 + O2 reaction has been measured.22 In their JSR measurements of 2M2B oxidation in the temperature range 600–1100 K and at about 1 atm pressure with residence time 1.5 s, Westbrook et al.16 observed only a little reactivity below 800 K, above which a rapid increase in the fuel consumption was observed. Under LTC conditions (500 < T < 1000 K), one may typically expect a fuel-radical + O2 reaction to constitute the major sink reaction of fuel-radical until unimolecular thermal decomposition reactions become accessible. However, in case of allyl-type radicals this is not necessarily the case (as discussed above for the CH3CHCHCH2 + O2 reaction), and for this reason kinetic oxidation experiments with still larger C5 allyl-type radicals are of great importance.
In this work, we have studied the kinetics of the (CH3)2CCHCH2 (R) + O2 reaction (1) over a wide 238–660 K temperature range and at low pressures (0.2–5.7 Torr), as well as thermochemistry of the (CH3)2CCHCH2 + O2 (CH3)2CCHCH2OO reaction (1a,−1a) in a narrower temperature range using photoionization mass spectrometry.
As shown in Scheme (2), the resonance hybrid structure of the (CH3)2CCHCH2 radical is composed of two resonance structures. The association of O2 with either the terminal (t) or non-terminal (nt) radical site of the (CH3)2CCHCH2 radical generates ROO(t) and ROO(nt) adducts, respectively, conveniently denoted as (CH3)2CCHCH2OO in reaction (1a,−1a).
![]() | ||
Scheme 2 Resonance structures contributing to the resonance hybrid structure of the (CH3)2CCHCH2 radical. |
The kinetic measurements performed in this work were focused on providing accurate bimolecular rate coefficient and equilibrium constant data of reaction (1). Subsequently, high-level quantum chemistry calculations and master equation (ME) modeling of reaction (1) were performed to complement the measured bimolecular rate coefficient and thermochemistry data.
Photolysis of 1-bromo-3-methyl-2-butene or 1-chloro-3-methyl-2-butene precursor at 248 nm radiation was performed using MPB ASX-750 or Coherent COMPex 201F excimer laser, respectively, to achieve a homogeneous [(CH3)2CCHCH2] along the flow-reactor axis.
(CH3)2CCHCH2Br + hν248 nm → (CH3)2CCHCH2 + Br | (P1a) |
→ Other products | (P1b) |
(CH3)2CCHCH2Cl + hν248 nm → (CH3)2CCHCH2 + Cl | (P2a) |
→ Other products | (P2b) |
The measurements presented in this study were primarily conducted using the bromide precursor, while the chloride precursor, which has a higher BDE(C–Cl) compared to BDE(C–Br), was principally used to show that the kinetic results do not depend on the selected photolytic precursor. Due to the smaller absorption cross-section of (CH3)2CCHCH2Cl at 248 nm in comparison to (CH3)2CCHCH2Br, experiments using the chloride precursor required higher [(CH3)2CCHCH2Cl] and higher laser fluences [∼170 mJ cm−2] than measurements using the bromide precursor, where lower precursor concentrations and laser fluences [∼17–50 mJ cm−2] were used. We estimate that [(CH3)2CCHCH2]t=0 ≤ 3 × 1011 cm−3 in the kinetic measurements using either precursor.
Producing the same (CH3)2CCHCH2 radical by photolysing two different precursors, (CH3)2CCHCH2Br and (CH3)2CCHCH2Cl, with the same 248 nm energy photons enables to some extent testing a potential isomerization of nascent (CH3)2CCHCH2 radicals immediately after photolysis before their thermalization. Fig. 1 presents relative enthalpies (ΔH(298K)) of the various (CH3)2CCHCH2 radical isomers and the transition states (TS) connecting them as well as maximum internal energies of the radicals after 248 nm photolysis of (CH3)2CCHCH2Br and (CH3)2CCHCH2Cl precursors (full horizontal lines), all calculated at the MN15/cc-pVTZ level in this work. In a recent study by this group, Döntgen et al.24 investigated the kinetics of the CH2CHCHCH2CH3 + O2 reaction and utilized 248 nm photolysis of BrCH2CHCHCH2CH3 and ClCH2CHCHCH2CH3 precursors for radical production. They performed time-dependent density functional theory calculations for electronic excitation energies of the bromide and chloride precursors and concluded that the total vibrational excitation of the CH2CHCHCH2CH3 radical after photolysis is about 15.0 and 13.4 kcal mol−1 for the bromide and chloride precursors, respectively. Both these energies were significantly below the ∼37 kcal mol−1 energy required for the lowest-energy CH2CHCHCH2CH3 → CH3CHCHCHCH3 isomerization in the system, showing that in those experiments only one isomer was produced during photolysis. Considering the very close similarity between those and the current experiments, i.e. the same photolysis energy and especially very similar photolytic precursors (e.g. BrCH2CHCHCH2CH3vs. (CH3)2CCHCH2Br and ClCH2CHCHCH2CH3vs. (CH3)2CCHCH2Cl), we estimate that the post-photolysis vibrational excitation of the (CH3)2CCHCH2 radical is also about 15.0 and 13.4 kcal mol−1 for the bromide and chloride precursors, respectively. Both these values, which contain some uncertainty due to the above estimation, are significantly below the calculated energy of the lowest isomerization transition state, TS1 = 38.9 kcal mol−1, see Fig. 1. The experimental conditions and the results of the kinetic measurements using chlorine and bromine precursors are presented in Tables 2–4.
![]() | ||
Fig. 1 Relative enthalpies (ΔH(298K)) of the various (CH3)2CCHCH2 radical isomers and the connecting transition states (TS) calculated at the MN15/cc-pVTZ level. The colored horizontal lines represent the difference between the 248 nm photon energy and the calculated BDE of Br–C and Cl–C bonds for (CH3)2CCHCH2Br (BDE(MN15/cc-pVTZ) = 55.1 kcal mol−1) and (CH3)2CCHCH2Cl (BDE(MN15/cc-pVTZ) = 68.4 kcal mol−1) precursors, respectively. The red, shaded area shows our best estimate on excitation energy of the (CH3)2CCHCH2 radicals after photolysis of (CH3)2CCHCH2Br and (CH3)2CCHCH2Cl precursors at 248 nm wavelength, respectively. The optimized geometries and ΔH(298K) values of the stationary points are given in the ESI† (see Fig. S1). |
The reaction mixture flowing through the temperature-controlled flow-reactor consisted of a mainly inert bath-gas (Helium > 97%) along with radical precursor ((CH3)2CCHCH2Br < 0.02%, (CH3)2CCHCH2Cl < 0.05%) and molecular oxygen ([O2] < 3%). In the high-temperature measurements a significantly higher amount of O2 (<51%) was used. Depending upon the temperature and pressure conditions employed during the measurements, different combinations of reactor tube sizes and coatings were used (see also Tables 2–4 caption): Pyrex-glass reactor (i.d. = 1.65 cm) with polydimethylsiloxane (PDMS) or halocarbon wax (HW) coating, stainless steel reactor (i.d. = 1.7 cm or 0.8 cm) with HW coating, and quartz reactor (i.d. = 0.85 cm) with boric oxide coating. A set temperature of the reaction zone was maintained within a uniformity of ±2 K using a PID controller.
The observed kinetics of the (CH3)2CCHCH2 + O2 reaction (1) at different temperatures can be characterized as follows.
![]() | (1a) |
![]() | (1a, −1a) |
![]() | (1b) |
Here, reaction (1a) represents the association reaction between the (CH3)2CCHCH2 radical and O2 leading to the formation of (CH3)2CCHCH2OO adducts. This reaction is important at low temperatures, 238–298 K. Reactions (1a,−1a) are important in the intermediate temperature range (325–373 K), where adducts become thermally unstable and a kinetic equilibrium is established between the forward bimolecular (CH3)2CCHCH2 + O2 reaction and the unimolecular dissociation of (CH3)2CCHCH2OO adducts back to the reactants. Phenomenological reaction (1b) corresponds to new reaction channels that open for (CH3)2CCHCH2 + O2 reaction at high temperatures (>500 K).
Apart from the reaction with O2, two other reactions that contribute to the (CH3)2CCHCH2 radical decay under experimental conditions are heterogeneous wall losses and second order radical-radical decay processes.
![]() | (A) |
R + R → Products | (B) |
The detection of the (CH3)2CCHCH2 radical decay profiles for reaction (1) was performed using the PIMS. The radiation of energy 8.44 eV from a microwave-discharge of 6% xenon in a helium-mixture with a sapphire window to cut-off higher energy photons was used to ionize (CH3)2CCHCH2 radicals present in the reaction mixture. Potential products of reaction (1b) were searched utilizing ionization radiation from a H2-lamp (discharge of H2 gas with an MgF2 window to produce radiation at around 10.2 eV). Ions formed were mass selected by a quadrupole mass spectrometer followed by the detection using an electron multiplier. Radical ion (CH3)2CCHCH2+ was detected at its parent mass at m/z = 69. A temporal ion decay signal was recorded from 10 ms before to 50 ms after each laser pulse, and accumulated for 3000–15000 repetitions before transferring to a PC for further analysis.
A nonlinear least-squares-fitting method was used for fitting the observed (CH3)2CCHCH2 radical decay profiles. For reactions (1a) and (1b) the measured radical decay profiles were fitted using the single-exponential expression: . Here,
is directly related to the bimolecular rate coefficients of the reaction (1a; k1a) and (1b; k1b) via relation:
B is the background signal, and [R0] and [Rt] are proportional to the initial (t = 0) and temporal radical concentration, respectively. Under conditions where kinetic equilibrium is established, a decay profile of a resonantly stabilized radical with two non-similar resonance structures follows, in principle, a tri-exponential behavior. However, following the scheme proposed by Knyazev et al.22 for the kinetics of the CH3CHCHCH2 + O2 reaction, such tri-exponential behavior can be fitted with a bi-exponential expression, [Rt] = B + Ce−λ1t + De−λ2t, which effectively leads to averaging the kinetic behavior of the two radical sites. Subsequently, the obtained fitted parameters (i.e. C, D, λ1, λ2) are used to calculate the forward (k1a) and backward (k−1a) rate coefficients of reaction (1a,−1a) using the following equations.
![]() | (E1) |
k−1a = λ1 + λ2 − (k1a[O2]) − k(R)wall − k(ROO)irr. | (E2) |
Here, k(ROO)irr. represents the rate coefficients for irreversible first-order decay processes (including heterogeneous wall losses) related to ROO adducts, and is calculated using expression (E3).
![]() | (E3) |
The values for equilibrium-constants (Keq) of reaction (1a,−1a) were calculated using expression (E4).
![]() | (E4) |
Here, the standard pressure term (P°) is introduced to the (E4) for satisfying the definition of equilibrium-constant being a dimensionless quantity. The species involved are assumed to have an ideal gas behavior and their standard states are chosen at 1 bar pressure (i.e. P° = 1 bar), at a specified temperature. R is the gas constant and NA is the Avogadro's constant.
Before use, 1-bromo-3-methyl-2-butene (Sigma-Aldrich, purity >95%) and 1-chloro-3-methyl-2-butene (Sigma-Aldrich, purity >95%) were degassed by several freeze–pump–thaw cycles. Helium (Messer-Griesheim purity of 99.9996%) and O2 (Aga, purity of 99.998%) were used as supplied.
EHF(X) = ECBSHF + BHF e−aHFX | (E5) |
Ecorr(X) = ECBScorr + BcorrXcorr−a | (E6) |
Here X is 2, 3, and 4 for the basis sets cc-pVDZ, cc-pVTZ, and cc-pVQZ, respectively and the ECBS, B, and a parameters are obtained by solving a system of three equations. The MN15/Def2TZVP and ROHF-CCSD(T)/CBS calculations were performed with Gaussian 16.34
Multireference methods were used to evaluate the energies of some stationary points. We used the CASPT2 method with the cc-pVTZ and cc-pVQZ basis sets. For multireference calculations, we extrapolated the complete basis set (ΔECBStotal) limit energy difference from the following relation.35
ΔECBStotal = ΔEcc-pVQZ − ((ΔEcc-pVTZ − ΔEcc-pVQZ) × 0.69) | (E7) |
The ORCA software package (program version 4.2.0)36 was employed in these calculations using the default CASSCF/CASPT2 settings. Notably, we did not use IPEA or level shifts. In a recent publication, Zobel et al. questioned the validity of the IPEA shift.37 In their investigation of excitation energies of organic molecules, they found that the ideal value of the IPEA shift correlated both with the size of the molecule and the basis set that was being used. Thus, the idea of there being an average IPEA shift value that improves the results for most systems seems unreasonable. Furthermore, for the benchmark set they were testing, they found that using the standard IPEA shift of 0.25 resulted in poorer agreement with experiment than not using an IPEA shift. Level shifts are typically used to avoid problems with intruder states. In our calculations, reference weights were between 0.7–0.8, so intruder states should not be a problem and, therefore, level shifts are not needed. Details on active space selections are given in the results and discussion section.
In order to complement the measured bimolecular rate coefficients (k1a, k1b), ME simulations were performed on the associated PES. ME simulations were done using MESMER code (version 6.1)38 incorporating 1-D energy-grained ME treatment. Coupling between external and internal rotations can be treated classically, although potential coupling between internal rotors is not explicitly treated.
Since there are no well-defined transition states for the barrierless R + O2 → ROO association reactions,24 we used the inverse-Laplace transform (ILT) method implemented in MESMER to compute the number of states of the variational transition states. The expression MESMER transforms is the modified Arrhenius expression of the canonical (high-pressure) association reaction rate coefficient. The Arrhenius parameters in the expression can be optimized using experimental data.
![]() | (E8) |
Here, k∞(T) is the rate coefficient at the high pressure limit, T is temperature, and Tref (=300 K) is the reference temperature. A, m and Ea are the pre-exponential factor, modified Arrhenius parameter and activation energy, respectively. Note that since reaction (1a) is considered to be a barrierless reaction, the activation energy in (E8) is approximated to be zero (i.e. Ea = 0). Microcanonical rate coefficients are calculated using RRKM theory.
MESMER utilizes a single-exponential-down model (E9) to account for the energy transfer via collisions between bath gas (i.e. helium in our case) and modelled species.
![]() | (E9) |
Here, 〈ΔEdown〉 corresponds to an average energy transferred from a higher energy grain to a lower energy grain in a single collision, and is used to calculate the associated collisional energy transfer probabilities. The 〈ΔEdown,ref〉 and n are the independent parameters representing the value of 〈ΔEdown〉 at some reference temperature (Tref = 300 K) and the temperature dependence of the 〈ΔEdown〉, respectively. The Lennard-Jones (L-J) potential model was used for obtaining the collision frequencies governing the energetics of collisions between bath-gas and modelled species. The L-J parameters used for helium bath-gas and the modeled species are shown in Table 1.
Species | ε (K) | σ (Å) |
---|---|---|
Helium | 10.2 | 2.55 |
C5H9O2 intermediates | 463 | 6.53 |
We used the online resources of Cantherm40 to estimate the L-J parameters of (CH3)2CCHCH2OOH (2-methylbut-3-ene-2-peroxol) and HOO(CH3)2CCHCH2 (3-methylbut-2-ene-1-peroxol) with the Joback method. We averaged these values and assigned them to the C5H9O2 intermediates.
An energy grain size of 75 cm−1 was used in all MESMER simulations. The quantum-mechanical tunneling corrections for elementary steps involving intra-molecular H-atom transfer were modelled using Eckart potential energy barriers.
Subsequently, using the experimental data as input, MESMER optimizes the fitting parameters for the modified Arrhenius eqn (E8) (i.e. parameters A and m) and the collisional energy transfer model (i.e. parameters 〈ΔEdown,ref〉 and n), by minimizing the value of χ2(k1a(exp.) − k1a(calc.))2 using a built-in Levenberg–Marquardt fitting algorithm (non-linear least-squares fitting method). The optimized ME model may then be used to extrapolate the kinetics of reaction (1) to experimentally inaccessible temperature and pressure conditions.
A typical bimolecular plot for reaction (1a) in the low temperature range is shown in Fig. 3 and the plot represents the measured pseudo-first-order rate coefficients as a function of the employed molecular oxygen concentration, [O2].
![]() | ||
Fig. 3 A bimolecular plot of the (CH3)2CCHCH2 + O2 → (CH3)2CCHCH2OO association reaction (1a) measured at T = 298 K and P = 1.2 Torr. A typical (CH3)2CCHCH2 radical decay profile (![]() |
We were not able to observe a (CH3)2CCHCH2OO adduct (m/z = 101) formation signal, because ions of peroxy radicals (apart from CH3OO radical) are generally not stable and may instead dissociatively ionize to produce the parent radical ion (i.e. (CH3)2CCHCH2+ radical ion in the current case) among other products.41 For this reason, in our kinetic experiments we have utilized a low photon energy, 8.44 eV, radiation for ionization of (CH3)2CCHCH2 radical (m/z = 69) and avoided any dissociative ionization of (CH3)2CCHCH2OO adducts to the m/z = 69 of the parent radical. This can be clearly observed from Fig. 2, where the signal at m/z = 69 from the (CH3)2CCHCH2 radical measured at 298 K decays back to the baseline.
For the low temperature range experiments, the measured pressure-dependent bimolecular rate coefficients (k1a) of reaction (1a) are shown in Table 2. Experiments were mainly performed using the (CH3)2CCHCH2Br photolytic precursor. A few measurements using chloride precursor show that the measured k1a values are independent of the precursor used (i.e. (CH3)2CCHCH2Br or (CH3)2CCHCH2Cl).
T (K) | P He (Torr) | [He]/1016 (cm−3) | [Pr]/1012 (cm−3) | [O2]/1013 (cm−3) |
k
(R)wall![]() |
k
(R)wall![]() |
k 1a/10−12 (cm3 molecule−1 s−1) | |
---|---|---|---|---|---|---|---|---|
a Xe-lamp with a sapphire window was used for detection as well as stainless steel reactor (i.d. = 1.7 cm) with halocarbon wax coating, unless otherwise stated. b Average of the measured wall rates for the (CH3)2CCHCH2 radical. c Wall rates determined from the linear fit y-axis intercepts of the bimolecular plots for reaction (1a). d Stainless steel reactor (i.d. = 0.8 cm) with halocarbon wax coating. e 1-Chloro-3-methyl-2-butene precursor was used for the production of (CH3)2CCHCH2 radical utilizing pyrex-glass reactor (i.d. = 1.65 cm) with halocarbon wax coating. | ||||||||
238 | 0.22 | 0.89 | 0.30 | 2.32–6.81 | 47.9–119.2 | 5.5 ± 0.8 | 5.5 ± 1.5 | 1.70 ± 0.09 |
238 | 0.40 | 1.62 | 0.27 | 1.45–4.65 | 38.2–94.3 | 5.1 ± 0.9 | 5.5 ± 1.1 | 2.06 ± 0.08 |
238 | 0.81 | 3.29 | 0.25 | 0.95–3.01 | 33.2–79.3 | 4.0 ± 0.6 | 4.2 ± 0.7 | 2.56 ± 0.09 |
238d | 2.39 | 9.70 | 1.87 | 2.68–5.13 | 109.6–167.8 | 19.7 ± 1.2 | 19.8 ± 0.9 | 2.97 ± 0.13 |
238d | 3.66 | 14.85 | 2.99 | 2.31–5.33 | 120.3–202.1 | 21.8 ± 1.2 | 22.4 ± 2.7 | 3.55 ± 0.16 |
268 | 0.20 | 0.72 | 0.25 | 4.11–15.06 | 43.6–129.0 | 4.1 ± 0.8 | 4.3 ± 0.8 | 0.85 ± 0.03 |
268 | 0.40 | 1.44 | 0.40 | 1.63–6.82 | 19.5–76.1 | 2.2 ± 0.9 | 2.1 ± 0.4 | 1.07 ± 0.02 |
268 | 0.81 | 2.92 | 0.39 | 1.11–4.96 | 25.2–79.9 | 3.5 ± 0.7 | 3.5 ± 0.7 | 1.54 ± 0.06 |
268e | 1.21 | 4.36 | 19.4 | 4.84–10.46 | 80.3–184.0 | 8.8 ± 2.6 | 8.5 ± 2.1 | 1.62 ± 0.09 |
268d | 2.68 | 9.66 | 3.03 | 2.16–8.36 | 69.5–185.7 | 19.8 ± 1.2 | 21.9 ± 3.6 | 1.97 ± 0.11 |
268d | 4.12 | 14.85 | 3.06 | 2.55–6.06 | 86.0–160.3 | 24.3 ± 1.3 | 24.2 ± 3.0 | 2.21 ± 0.12 |
298 | 0.31 | 1.00 | 0.35 | 4.28–12.52 | 17.1–60.2 | 2.8 ± 0.7 | 2.5 ± 0.8 | 0.45 ± 0.02 |
298 | 0.60 | 1.94 | 0.32 | 2.82–12.65 | 24.0–91.7 | 4.4 ± 0.7 | 4.4 ± 0.7 | 0.67 ± 0.02 |
298 | 1.20 | 3.89 | 0.32 | 3.11–12.85 | 33.0–132.1 | 2.5 ± 0.6 | 2.4 ± 0.6 | 0.97 ± 0.02 |
298e | 1.20 | 3.89 | 16.9 | 4.55–12.28 | 53.4–134.2 | 3.7 ± 1.9 | 3.8 ± 0.7 | 1.02 ± 0.02 |
298d | 2.99 | 9.69 | 0.84 | 3.63–25.91 | 74.9–297.6 | 17.4 ± 1.0 | 18.4 ± 2.2 | 1.19 ± 0.08 |
298d | 4.57 | 14.81 | 1.68 | 4.25–11.33 | 81.1–164.7 | 15.8 ± 0.7 | 16.4 ± 2.1 | 1.40 ± 0.06 |
For the intermediate temperature range, the observed forward and backward rate coefficients of reactions (1a) and (−1a), along with respective natural logarithm of the equilibrium constants (ln(Keq)) are given in Table 3. The results of the intermediate temperature range measurements are also independent of the precursor ((CH3)2CCHCH2Br or (CH3)2CCHCH2Cl) used. The returned values of the fitting parameters associated with fitting the bi-exponential expression [Rt] = B + Ce−λ1t + De−λ2t to the decay traces are given in the ESI† (see Table S1).
T (K) | P He (Torr) | [He]/1016 (cm−3) | [O2]/1014 (cm−3) |
k
(R)wall![]() |
k
(ROO)irr.![]() |
k 1a/10−14 (cm3 molecule−1 s−1) | k −1a (s−1) | ln(Keq)d |
---|---|---|---|---|---|---|---|---|
a Xe-lamp with a sapphire window was used for detection as well as Pyrex-glass reactor (i.d. = 1.65 cm) with polydimethylsiloxane (PDMS) coating, unless otherwise stated. b Average of the measured wall rates for the (CH3)2CCHCH2 radical. c Rate coefficients for irreversible first-order decay processes of ROO adducts, determined from the bi-exponential fit performed to the observed (CH3)2CCHCH2 radical decay profiles. d Natural logarithmic of the measured Keq values. e 1-Chloro-3-methyl-2-butene precursor was used for the production of the (CH3)2CCHCH2 radical as well as Pyrex-glass reactor (i.d. = 1.65 cm) with halocarbon wax coating. f Stainless steel reactor (i.d. = 0.8 cm) with halocarbon wax coating. | ||||||||
325 | 0.86 | 2.56 | 2.99 | 13.8 ± 0.3 | 15.4 ± 1.5 | 48.7 ± 6.4 | 15.0 ± 4.7 | 13.49 ± 0.34 |
325 | 0.86 | 2.56 | 4.83 | 14.0 ± 0.3 | 18.1 ± 1.8 | 44.8 ± 5.9 | 15.6 ± 6.4 | 13.37 ± 0.43 |
330 | 0.88 | 2.58 | 5.37 | 14.6 ± 0.3 | 18.8 ± 2.2 | 43.0 ± 5.6 | 21.7 ± 10.3 | 12.98 ± 0.49 |
335 | 0.90 | 2.59 | 4.36 | 12.3 ± 0.2 | 11.8 ± 1.1 | 35.8 ± 4.5 | 24.4 ± 5.2 | 12.67 ± 0.25 |
337 | 0.90 | 2.58 | 4.88 | 13.5 ± 0.3 | 15.7 ± 1.4 | 35.1 ± 4.5 | 33.3 ± 8.2 | 12.33 ± 0.28 |
338 | 0.91 | 2.60 | 5.17 | 17.7 ± 0.3 | 17.2 ± 1.5 | 34.1 ± 4.3 | 31.5 ± 7.8 | 12.35 ± 0.28 |
344 | 0.92 | 2.58 | 5.01 | 15.5 ± 0.3 | 17.7 ± 1.1 | 31.1 ± 3.9 | 49.6 ± 7.5 | 11.79 ± 0.20 |
348 | 0.94 | 2.61 | 5.35 | 15.0 ± 0.5 | 16.5 ± 1.1 | 26.0 ± 3.2 | 56.7 ± 7.8 | 11.47 ± 0.19 |
353 | 0.96 | 2.63 | 5.73 | 13.2 ± 0.5 | 20.0 ± 4.4 | 19.5 ± 3.3 | 63.2 ± 25.7 | 11.05 ± 0.44 |
356 | 0.96 | 2.60 | 5.22 | 12.9 ± 0.4 | 19.5 ± 1.8 | 21.8 ± 2.8 | 98.5 ± 13.9 | 10.71 ± 0.19 |
356e | 1.22 | 3.31 | 7.39 | 9.8 ± 2.1 | 7.4 ± 5.1 | 24.4 ± 11.3 | 100.8 ± 76.9 | 10.80 ± 1.09 |
358 | 0.97 | 2.62 | 6.37 | 14.6 ± 0.5 | 15.1 ± 5.9 | 21.7 ± 5.7 | 116.6 ± 66.7 | 10.54 ± 0.63 |
361 | 0.48 | 1.28 | 5.58 | 12.6 ± 0.3 | 20.5 ± 2.2 | 15.6 ± 3.9 | 79.5 ± 12.7 | 10.58 ± 0.30 |
361 | 0.96 | 2.57 | 7.17 | 13.8 ± 0.3 | 16.1 ± 1.5 | 21.3 ± 2.8 | 124.8 ± 17.8 | 10.44 ± 0.19 |
361e | 1.22 | 3.26 | 7.21 | 9.9 ± 2.0 | 7.9 ± 4.8 | 18.2 ± 8.1 | 116.1 ± 85.4 | 10.36 ± 1.02 |
361f | 3.61 | 9.66 | 6.57 | 15.9 ± 0.9 | 33.1 ± 14.9 | 13.5 ± 4.1 | 103.4 ± 67.8 | 10.17 ± 0.72 |
361f | 5.58 | 14.93 | 4.72 | 17.5 ± 1.1 | 39.6 ± 21.2 | 19.5 ± 7.1 | 128.5 ± 95.0 | 10.32 ± 0.82 |
362 | 0.97 | 2.59 | 5.20 | 12.6 ± 0.3 | 21.4 ± 2.8 | 22.3 ± 3.3 | 153.7 ± 25.7 | 10.27 ± 0.22 |
367 | 0.71 | 1.87 | 6.45 | 13.3 ± 0.5 | 19.2 ± 4.4 | 17.7 ± 3.9 | 167.7 ± 44.6 | 9.95 ± 0.35 |
373 | 0.76 | 1.97 | 4.66 | 14.2 ± 0.4 | 19.2 ± 5.9 | 21.7 ± 5.7 | 264.8 ± 71.0 | 9.67 ± 0.38 |
373 | 1.21 | 3.13 | 6.48 | 12.4 ± 0.5 | 20.8 ± 6.0 | 17.5 ± 3.9 | 282.2 ± 75.3 | 9.40 ± 0.35 |
The stated uncertainties of the k1a values in the low temperature range (Table 2) originate from linear fits to the data and are somewhat smaller than the uncertainties of the k1a values in the intermediate temperature range (Table 3), where the bi-exponential expression described above is fitted to the data and the stated uncertainties originate from using an error propagation method and eqn (E1). The stated uncertainties of the k−1a values in the intermediate temperature range (Table 3) are obtained using an error propagation method and eqn (E2). The uncertainties of the k−1a values are larger than those of the k1a values.
The overall uncertainty (accuracy) of the k1a values presented in Table 2 is estimated to be about ±25%. This estimation is composed of several sources of uncertainties. The above discussed fitting uncertainties are the single most important sources, while uncertainties to determine reactant concentrations and uncertainties in measuring temperature and pressure of the reaction zone also play a role. Overall uncertainties of the k1a and k−1a values in the intermediate temperature range are difficult to estimate.
![]() | ||
Fig. 4 A bimolecular plot for (CH3)2CCHCH2 + O2 → products reaction (1b), measured at T = 550 K and P = 2.92 Torr. The lower right corner of the plot presents the (CH3)2CCHCH2 radical decay profile ![]() |
T (K) | P He (Torr) | [He]/1016 (cm−3) | [Pr]/1012 (cm−3) | [O2]/1015 (cm−3) |
k
(R)wall![]() |
k
(R)wall![]() |
k 1b/10−15 (cm3 molecule−1 s−1) | |
---|---|---|---|---|---|---|---|---|
a Xe-lamp with a sapphire window was used for detection as well as quartz reactor (i.d. = 0.85 cm) with boric acid coating. b Average of the measured wall rates for the (CH3)2CCHCH2 radical. c Wall rates determined from the linear fit y-axis intercepts of the bimolecular plots for reaction (1b). | ||||||||
501 | 2.90 | 5.59 | 1.25 | 7.41–29.17 | 72.8–120.1 | 29.6 ± 2.3 | 33.3 ± 4.1 | 2.82 ± 0.44 |
502 | 2.94 | 5.66 | 2.71 | 9.84–47.43 | 82.1–185.5 | 44.6 ± 1.9 | 45.0 ± 1.2 | 3.08 ± 0.14 |
550 | 2.92 | 5.13 | 1.31 | 6.66–28.32 | 46.6–141.3 | 19.7 ± 1.7 | 18.7 ± 2.7 | 4.23 ± 0.35 |
579 | 1.83 | 3.05 | 2.48 | 3.95–7.52 | 20.6–33.6 | 5.2 ± 0.4 | 5.22 ± 0.1 | 3.75 ± 0.10 |
660 | 5.67 | 8.30 | 4.38 | 7.26–23.28 | 56.8–103.3 | 9.1 ± 1.2 | 9.5 ± 1.9 | 4.62 ± 0.43 |
Measuring kinetics as slow as reaction (1b) with k1b ≈ 3–5 × 10−15 cm3 molecule−1 s−1 in this work requires significant amounts of O2 (<51% of the total reaction mixture) to be added to the reaction mixture. This results in high total pressure in the reaction zone (Ptotal < 7.8 Torr) and high oxygen content of the reaction mixture, which altogether causes high pressures in the vacuum chamber containing the quadrupole mass spectrometer (PQMS-chamber). In fact, conditions (i.e. PQMS-chamber) of the kinetic measurements of reaction (1b) were at or even beyond the upper limit of pressure that our PIMS setup can tolerate. In addition, at T > 500 K, presumably heterogeneous wall-effects started to have an influence on the measured values, causing some additional uncertainty on the respective k1b values. There is some more discussion on this in the ESI† (see Fig. S2). Taking into account the abovementioned additional uncertainty sources at T > 500 K, including the uncertainties in the temperature and pressure measurements in the reaction zone as well as the fitting uncertainties, the overall uncertainty (=accuracy) in the k1b values presented in Table 4 is estimated to be the factor of two (upper = 2 × k1b, lower = 0.5 × k1b).
Measurements were also conducted to detect possible products of reaction (1b). Our quantum chemistry calculations (see next section), suggest isoprene (m/z = 68) + HO2 (m/z = 33) to be the major product channel of reaction (1b). Our PIMS apparatus is unable to detect HO2 radical, mainly due to its high ionization energy IE(HO2) ≈ 11.4 eV.42 Utilizing H2-lamp for ionization radiation, a product formation signal was observed at m/z = 68, see Fig. 5. However, a fast, almost immediate product formation signal at m/z = 68 was observed also in the absence of [O2], originating, almost certainly, from the photolysis of the radical precursor (CH3)2CCHCH2Br. For this reason, two consecutive measurements were performed, the first one in the absence and the second one in the presence of O2, see Fig. 5(a). The signal without O2 was then subtracted from the signal with O2 present, giving a product formation signal at m/z = 68 originating only from reaction (1b), see Fig. 5(b). The product formation signal at m/z = 68 was fitted using the expression where, It(P+) is the temporal product ion signal and I0(P+) the signal value for the channel yield × [R]0 as t → ∞. The term
was treated as an independent parameter and the fitting process returned the value
see the fit in Fig. 5(b). This value agrees well within the stated 1σ-uncertainty with the
value obtained from fitting the corresponding (CH3)2CCHCH2 radical decay profile shown in Fig. 5(c). This shows that under utilized pseudo-first-order conditions, signal at m/z = 68 originates from reaction (1b). As already mentioned, high values of PQMS-chamber reduces detection sensitivity of the PIMS apparatus, which is evident from the product ion signals presented in the Fig. 5.
![]() | ||
Fig. 5 The product formation and corresponding radical decay profiles for the (CH3)2CCHCH2 + O2 → products reaction (1b), measured at T = 501 K, P = 2.9 Torr. The top most panel (a) depicts the product ion formation signals observed at m/z = 68. The filled blue and red circles represent the product ion formation signals when [O2] = 0 cm−3 and 2.2 × 1016 cm−3, respectively. The middle panel (b) shows the difference between the product ion formation signals observed at [O2] = 2.2 × 1016 cm−3 and [O2] = 0 cm−3. The corresponding (CH3)2CCHCH2 radical decay profile measured at [O2] = 2.2 × 1016 cm−3![]() |
The measured values of k1b exhibit weak, positive temperature dependence. Current results show that in the temperature range of 500–660 K, the k1b values are in the range of (2.82–4.62) × 10−15 cm3 molecule−1 s−1. These values are comparable to the bimolecular rate coefficient k(753 K) = 3.49 × 10−15 cm3 molecule−1 s−1 measured by Baldwin et al.23 for the kinetics of the CH2CHCHCH2CH3 + O2 → CH2 = CHCH = CHCH3 + HO2 reaction channel at 753 K, and about 10 times faster than an upper-limit of bimolecular rate coefficient of the CH3CHCHCH2 + O2 reaction measured at 700 K by Knyazev et al.22 In addition, current results are an order of 103 times faster than the bimolecular rate coefficient measured for allyl-radical + O2 reaction at a temperature of 673 K.20 We are not aware of any previous measurements of bimolecular rate coefficients of the (CH3)2CCHCH2 + O2 reaction.
The Arrhenius plot for the forward bimolecular rate coefficients of the (CH3)2CCHCH2 + O2 reaction, k1a and k1b, measured in this work at the overall temperature range of (238–660) K, is presented in Fig. 6. This plot is both striking and revealing. It shows that in both the low- and intermediate temperature ranges the (CH3)2CCHCH2 + O2 reaction shows negative temperature and positive pressure dependence and is faster than the allyl + O2 reaction19 but comparable to the CH2CHCHCH2CH3 + O2 reaction24 at 298 K. Upon increasing temperature to about 500 K, the rate coefficient of the (CH3)2CCHCH2 + O2 reaction decreases by almost two orders of magnitude. The temperature dependence also changes from negative to positive and the rate coefficient becomes pressure-independent. Next, we use computational methods to better understand the kinetics and mechanism behind the observed behavior of the (CH3)2CCHCH2 + O2 reaction.
We observe that CASPT2 calculations systematically predict lower barriers in comparison to ROHF-CCSD(T) calculations. Significant reductions in barrier heights are observed even in cases where the T1 diagnostic suggest that a single-reference method is suitable. The obvious question is which energies should we trust? The ROHF-CCSD(T) method produces more reliable results when a single-reference approach is suitable, and for the many important transition state structures the T1 diagnostics suggest that such an approach is suitable. However, we are confident in our experimental findings, which strongly suggest that the ROHF-CCSD(T) energies (barriers) are too high and that the CASTP2 energies are more reasonable. If the CASPT2 energies are indeed more correct, it may indicate that the T1 diagnostic alone is a less reliable measure of the multi-reference character present in the studied species. Although we are more inclined to trust the CASPT2 energies, we cannot entirely rule out the possibility that the ROHF-CCSD(T) energies are reliable, in which case we have simply failed to locate low-energy reaction channel(s). For ME simulations, we decided to use ROHF-CCSD(T) energies for the peroxyl radicals (ROO(t), ROO(nt) well-depths) and for the products. CASPT2 energies are used for everything else. To simplify the simulations, we left out all reaction channels with a barrier higher than 8 kcal mol−1 above the energy of the reactants.
In our modeling effort, we first used our experimental low- and intermediate-temperature kinetic data to obtain the optimal values for the ROO well-depths as well as for both the exponential-down collision energy transfer (E9) and modified Arrhenius expression (E8) parameters. For the barrierless association reaction channels R + O2 ⇌ ROO(t) and R + O2 ⇌ ROO(nt), MESMER returned the following modified Arrhenius expressions at the high-pressure limit (Ea in (E8) was set = 0): and
, respectively. The optimized 〈ΔEdown〉 parameters are presented by the following expression:
. See also Table 5 for these values. For helium bath-gas, the obtained 〈ΔEdown〉 parameters (i.e. 〈ΔEdown,ref〉 = 166 cm−1, n = 0.1), lies within their typical limits.38
〈ΔEdown,ref〉 (cm−1) | n | A R + O2 ⇌ ROO(t) (cm3 molecule−1 s−1) | A R + O2 ⇌ ROO(nt) (cm3 molecule−1 s−1) | m R + O2 ⇌ ROO(t) | m R + O2 ⇌ ROO(nt) |
---|---|---|---|---|---|
166 | 0.1 | 1.17 × 10−12 | 2.34 × 10−12 | (−1.06) | (−1.04) |
In the fitting process, the energy difference between ROO(nt) and ROO(t) well-depths was constrained to remain equal to the energy difference obtained from ROHF-CCSD(T)/CBS calculations. The fitted ROO well-depths are 0.72 kcal mol−1 shallower than the computed ones, but are within estimated computational uncertainty (∼1 kcal mol−1).
As shown in Fig. 7, the measured values of average ln(Keq) at the intermediate temperature range are plotted alongside the calculated channel-specific ln(Keq) values of reaction channels R + O2 ⇌ ROO(t) and R + O2 ⇌ ROO(nt). For a given temperature range, MESMER uses NASA-polynomials for generating the temperature dependent standard thermodynamic quantities, such as standard enthalpy (H°)T, entropy (S°)T and Gibbs free energy (G°)T, for different reaction species. These standard thermodynamic quantities are then used for calculating the ln(Keq) values for each reaction channel independently. The thermodynamic MESMER output file contains the standard thermodynamic quantities of a particular species in their zero-point energy corrected form e.g.. This enables us to locate the changes in the calculated channel specific ln(Keq) values with respect to the fitted ROO well-depths. Interestingly, the more shallow ROO(nt) and ROO(t) well-depths (by 0.72 kcal mol−1 each) returned by the fitting process, give the ln(Keq) values for reaction channel R + O2 ⇌ ROO(nt) that are equivalent to the measured average ln(Keq) values, see Fig. 7. This strongly suggests that the kinetic traces measured in the intermediate temperature range are almost entirely due to the R + O2 ⇌ ROO(nt) reaction channel.
Incorporating 0.72 kcal mol−1 (shallower) correction to the ROO(nt) and ROO(t) well-depths provides the standard reaction enthalpies for reaction channels R + O2 ⇌ ROO(t) and R + O2 ⇌ ROO(nt) to be (−17.1) kcal mol−1 and (−20.0) kcal mol−1, respectively.
The potential energy profile utilized in the ME simulations of reaction (1), already including the 0.72 kcal mol−1 correction to the ROO well-depths, is shown in Fig. 8. The product energies are from ROHF-CCSD(T)/CBS calculations and all the other energies, apart from the ROO well-depths, are from CASPT2 calculations.
![]() | ||
Fig. 8 The simplified potential energy profile for the R + O2 ⇌ reaction (R = (CH3)2CCHCH2) used in our ME simulations. Energies are in kcal mol−1. The reported Int1T (ROO(t)) and Int1NT (ROO(nt)) energies include the correction of 0.72 kcal mol−1. The rest of the energies are either from ROHF-CCSD(T)/CBS//MN15/Def2TZVP or CASPT2/CBS//MN15/Def2TZVP calculations. The optimized geometries, T1 diagnostic values and single point energies of the stationary points are given in the ESI† (see Tables S2–S3). |
Measured bimolecular rate coefficients of reaction (1a) at three temperatures along with the results of constrained ME simulations performed over a wide pressure range are shown in Fig. 9. As can be seen, a reasonably good agreement with the experimental results is obtained.
![]() | ||
Fig. 9 The bimolecular rate coefficients of reaction (1a), R + O2 → ROO(t,nt), R = (CH3)2CCHCH2, plotted as a function of pressure at three temperatures. The symbols depict the experimental results and the lines show the results of ME simulations. The optimized modified Arrhenius equation parameters at the high-pressure limit and 〈ΔEdown〉 parameters are shown on the lower right corner. Filled and blank circles indicate the experiments conducted with brominated ((CH3)2CCHCH2Br) and chlorinated ((CH3)2CCHCH2Cl) precursors, respectively. |
While the constrained ME model is able to reproduce the experimental kinetic data in the low-temperature range, the simulated bimolecular rate coefficients in the high-temperature range are a factor of 5–9 times lower than the respective experimental k1b values. To remedy this, we lowered the barrier height of transition state (TS1NTP1) associated with the kinetically most dominant channel R + O2 → isoprene + HO2 by ∼2.2 kcal mol−1. This adjustment is within the expected uncertainty of the CASPTP2/CBS calculations.
During our ME simulations, we discovered that under most pressure conditions there is a one-to-one correspondence between an eigenvalue λ1 and the high-temperature bimolecular rate coefficients k1b at T > 450 K. This correspondence was checked at different temperatures and pressures by manually performing single-exponential fits to simulated decay profiles. Similarly, we also observed that an eigenvalue λ2 corresponds to the bimolecular rate coefficients measured in the low temperature range (T < 300 K). The pressure and temperature dependent behavior of λ1 and λ2 is illustrated in Fig. 10. Interestingly, under extremely low pressure conditions (P ∼ 1 × 10−7 bar) there is a nearly seamless transition from λ1 to λ2 upon increasing temperature. Above this pressure, there exists a narrow transition region (∼300–450 K) where the kinetic behavior of the system cannot be described by a single eigenvalue. In this region multi-dimensional decays are observed.
We multiply −λ1/[O2] with the calculated branching ratios of different product channels to obtain channel-specific bimolecular rate coefficients at T > 450 K. Arrhenius fits were then performed to obtain the results in a form that is usable in autoignition and combustion models. The channel specific Arrhenius fit parameters are reported in Table 6.
Reaction channel (T = 450–2000 K) | |||
---|---|---|---|
A (cm3 molecule−1 s−1) | m | E a (kcal mol−1) | |
TS1NTP1 lowered by ∼2.2 kcal mol−1 | |||
R + O2 → P1 | 1.46 × 10−19 | 1.71 | 0.78 |
Transition states unadjusted | |||
R + O2 → P1 | 2.87 × 10−20 | 1.91 | 2.46 |
R + O2 → P3 | 2.09 × 10−28 | 3.76 | −2.53 |
R + O2 → P4 | 3.56 × 10−18 | 0.79 | 7.02 |
R + O2 → P6 | 3.18 × 10−18 | 0.77 | 5.51 |
The Arrhenius expressions of Table 6 fit closely the results of ME simulations of bimolecular reaction channels leading to products P1, P3, P4, and P6, which are shown in Fig. 11(a). Among all the channels studied in this work, (CH3)2CCHCH2 + O2 → isoprene + HO2 (P1) is overwhelmingly the dominating one with a yield >91% or so without any adjustment to the stationary-point energies. This corroborates our assignment of product formation signal observed at m/z = 68 (see Fig. 5) originating from isoprene. The second most important channel is (2R/S)-3-methyl-1,2-epoxybut-3-ene + OH (P3) with yield <8%. In Fig. 11(b), the ME simulated k1b values obtained for different channels are compared to the current experimental k1b values. Clearly, an adjustment to stationary-point energy, here to TS1NTP1, is required to reach agreement with the experimental results. Also shown in Fig. 11(b) are the bimolecular rate coefficients of the (CH3)2CCHCH2 + O2 → isoprene + HO2 (P1) reaction channel used by Westbrook et al.16 in their kinetic modeling study of 2M2B oxidation. Apparently, the current ME simulated k1b values obtained for the respective (CH3)2CCHCH2 + O2 → P1 reaction channel only intersect with the results of Westbrook et al. at around T ∼ 1600 K, and diverge sharply for temperatures above and especially below 1600 K. The MESMER input file ((CH3)2CCHCH2.xml) used for ME modeling is available in the ESI.†
![]() | ||
Fig. 11 The graph (a) shows the results of ME simulations of different product channels of reaction k1b (solid lines) as well as their respective Arrhenius fits (dashed lines) given in Table 6. The graph (b) depicts the same results of ME simulations as in (a), compared with our experimental k1b values (symbols). The graph (b) also shows the rate-coefficients of the (CH3)2CCHCH2 + O2 → isoprene + HO2 (P1) bimolecular reaction channel estimated by Westbrook et al.16 in their 2M2B oxidation study. |
Probably the main outcome of the current, combined experimental and theoretical study of the allylic C5 radical reaction (CH3)2CCHCH2 + O2 is that larger (>C4) allylic radicals seem to be more reactive with respect to O2 under low-temperature combustion (LTC) conditions (T < 1000 K) than is currently believed. However, more work is certainly needed on oxidation kinetics of allylic radicals under LTC conditions to better understand their effects on autoignition behavior of fuel–air mixtures containing alkenes. Similar to this work, investigating oxidation kinetics of different allylic C5 (and larger) radical isomers experimentally (and theoretically) over a wide temperature range would be valuable. Also, it would be highly useful to have photolysis-initiated product studies of alkene oxidation experiments available that have been performed under LTC conditions.
Footnote |
† Electronic supplementary information (ESI) available: The optimized geometries and ΔH(298K) values of the stationary points associated with the isomerization of (CH3)2CCHCH2 radical, bi-exponential fitting parameters for intermediate temperature range measurements, determination of experimental uncertainty in the high-temperature range measurements, energies of the stationary points associated with reaction (1) along with their T1 diagnostic values and optimized geometries, and the MESMER input file ((CH3)2CCHCH2.xml), are provided in the electronic supplementary information. See DOI: 10.1039/d1cp02210e |
This journal is © the Owner Societies 2021 |