Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Article

Cite This: Energy Fuels 2019, 33, 11781−11794 pubs.acs.org/EF

Reaction of Hydroperoxy Radicals with Primary C1−5 Alcohols: A


Profound Effect on Ignition Delay Times
Saleh E. Rawadieh,†,‡ Ibrahem S. Altarawneh,§ Mohammad A. Batiha,† Leema A. Al-Makhadmeh,#
Mansour H. Almatarneh,∥ and Mohammednoor Altarawneh*,⊥

Chemical Engineering Department, ‡Renewable Energy Research and Development Center, and #Environmental Engineering,
Al-Hussein Bin Talal University, Ma’an 71111, Jordan
§
Pharmaceutical and Chemical Engineering Department, German Jordanian University, Amman 11180, Jordan

Department of Chemistry, University of Jordan, Amman 11942, Jordan
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Department of Chemical and Petroleum Engineering, United Arab Emirates University, Sheikh Khalifa bin Zayed Street, Al-Ain
15551, United Arab Emirates
Downloaded via UNITED ARAB EMIRATES UNIV on May 20, 2023 at 10:28:26 (UTC).

*
S Supporting Information

ABSTRACT: Ignition delay times of primary alcohols often display a noticeable sensitivity to their initial reactions with HO2
radicals. In view of the transient nature of HO2 radicals, kinetic models on combustion of alcohols utilize theoretically obtained
constant parameters for the abstraction HO2 + alcohols reactions. Rate constants for the title reactions in pertinent kinetic
models are often extrapolated from analogous computed values for either alkanes + HO2 or n-butane + HO2 reactions. Even for
the simplest alcohol, methanol, literature values for the reaction rate constants considerably vary within one order of magnitude.
Herein, we compute reaction rate constants for H abstraction from the weakest sites in primary C1−5 alcohols by HO2
(methanol, ethanol, n-propanol, i-propanol, n-butanol, i-butanol, t-butanol, n-pentanol, and i-pentanol). In most cases, our
reaction rate coefficients tend to slightly exceed corresponding values deployed in pertinent kinetic models. We have thoroughly
assessed the predictive performance of literature kinetic models in computing ignition delay times of these alcohols based on the
updated rate constants for HO2-abstraction reactions. In the case of methanol, updating kinetic parameters for the reaction
CH3OH + HO2 → CH2OH + H2O2 improves prediction of ignition delay times at lower temperatures in reference to original
literature kinetic models. Likewise, a modified kinetic model for n-butane and t-butanol affords better agreement with
experimental values of ignition delay times at low temperatures and high pressures. Kinetic parameters presented herein will be
useful to accurately account for salient oxidation features of alcohols in real combustion engines.

1. INTRODUCTION Reactions in Scheme 1 exhibit temperature- as well as


Hydroperoxy (HO2) radicals assume a key role in deriving the pressure-dependent behavior.7 HO2 radicals are mainly
low-temperature oxidation and auto-ignition of hydrocarbons consumed via H-abstraction reactions with hydrocarbons;
(typically between 600 and 1200 K).1 The commonly observed RH + HO2 → R· + H2O2. This reaction traps the relatively
“negative temperature coefficient, NTC” zone partly stems active HO2 radicals into the stable hydrogen peroxide
from the combustion cycle of HO2.2−4 Formation of HO2 in molecules (H2O2) and thus underpins the reduction in
combustion medium initially ensues via the endothermic and reactivity in the NTC region (around 650 K).3 As the
highly reversible O2-induced H abstraction reaction RH + 3O2 temperature increases, fragmentation of H2O2 into OH radicals
→ R· + HO2 where R· signifies a hydrocarbon radical. At restores the oxidation capacity. As Scheme 1 portrays, HO2
intermediate temperatures, reactions of hydrogen atoms with radicals chiefly participate in all stages pertinent to the
oxygen molecules produce an appreciable concentration of development of ignition and the low-temperature oxidation.
HO2. However, the HO2 radical predominantly forms via the Consequently, profiles of products and ignition delay times in
“cool-flame” reaction sequence depicted in Scheme 1 (RH oxidation models of various hydrocarbons were found to be
denotes an alkane species).5,6 highly sensitive to the kinetics of the abstraction reaction RH +
HO2 → R· + H2O2.8,9
Scheme 1. Cool-Flame Reaction Despite being highly transient species, sophisticated
experimental techniques provided direct identification and
quantification of HO2 radicals under atmospheric and
combustion conditions.10,11 Profiles of HO2 radicals were
qualitatively measured during oxidation of dimethylether12 and

Received: July 2, 2019


Revised: October 7, 2019
Published: October 14, 2019

© 2019 American Chemical Society 11781 DOI: 10.1021/acs.energyfuels.9b02169


Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

n-butane13 deploying optical approaches encompassing mid- relative contribution of HO2 radicals in deriving the low-
infrared Faraday rotation spectroscopy and fluorescence assay temperature oxidation of these alcohols is thoroughly assessed.
by gas expansion (FAGE), respectively. Nonetheless, acquired
absolute HO2 concentrations were deemed to be “uncertain” 2. COMPUTATIONAL METHODOLOGY
owing to sampling and calibration difficulties. Due to these The Gaussian 0934 suite of programs performs geometry
challenges and in view of the absence of a suitable HO2 optimizations, total energy computations, and vibrational
precursor, direct measurement of the reaction rate constant for frequency calculations at the composite chemistry model
H abstraction by HO2 is rather very limited. Jemi-Alade et al.14 CBS-QB3.35 Transition structures retain a single imaginary
measured the rate constant for the reaction HO2 + CH2O → frequency along the designated reaction coordinate. The CBS-
H2O2 + CHO in a flash photolysis experiment. Walker and co- QB3 methods locate optimized geometries at the B3LYP/
authors15,16 provided indirect measurements for H abstraction CBSB7 level of theory and employ successive single-point
from distinct C−H sites in alkanes, toluene, and benzene. Most energy calculations at higher theoretical levels. The satisfactory
kinetics data on the abstraction RH + HO2 reaction stem from performance of the cost-effective CBS-QB3 method in
quantum chemical calculations. For instance, theoretically predicting thermochemical and kinetic parameters for H
calculated reaction rate constants for H abstraction from abstraction reactions is very well established. In a recent
alkanes (up to butanes) by Aguilera-Iparraguirre et al.17 were paper,36 we compared reaction and activation energies
in accord with recommended analogous values by Walker.15 computed by the CBS-QB3 method (for H abstraction
Likewise, we have utilized density functional theory (DFT) reactions by NH2 radicals) against corresponding values
calculations and chemistry models to compute rate constants obtained with the more computation-intensive chemistry
for H abstraction by HO2 from a large array of compounds models, G4 and CBS-APNO. It was found that CBS-QB3
including methanol,18 benzene,19 mercaptans,20 and alkylated and CBS-APNO values reside within 2.0 kJ/mol. The mean
benzenes.21,22 Despite the progress on the theoretical side, the unsigned errors for the difference in activation enthalpies
literature presents a rather limited account of kinetics data between CBS-QB3 and G4 values incur a value of 3.0 kJ/mol.
germane to H abstraction by HO2 from an important fuel’s We also refer the reader to comprehensive thermochemistry
category, that is, alcohols. benchmarking of these three methods reported by Simmie et
Alcohols are important non-petroleum fuels on their own al.37,38 The CBS-QB3 method is typically executed at a
right or when blended with other hydrocarbon fuels. significantly shorter time than other chemistry models. For this
Combustion chemistry of these alternative fuels (mainly up reason, the CBS-QB3 method largely retains its position as the
to five carbon atoms) has been thoroughly investigated.23−27 most widely utilized cost-effective and accurate approach in
Laboratory- and engine-scale studies presented focal combus- computing thermokinetics parameters pertinent to combustion
tion properties of various types of alcohols with a prime focus reactions. The multiplicity of all initial abstraction reactions
on pyrolysis and oxidation mechanisms, flame propagation, amounts to 2.
and profiles of pollutant emission. Kinetic models on oxidation KiSThelP,39 a kinetic and statistical thermodynamic code,
of alcohols establish a noticeable sensitivity for the H computes reaction rate constants (fitted in the wide temper-
abstraction reaction by HO2.25,28 For instance, Sarathy et ature range 500−2000 K). It has been shown that variational
effects on the kinetics of HO2 + methanol/ethanol reactions
al.29 found that the ignition time of 1-butanol was the most
are rather very limited.31,40 Thus, reaction rate constants were
sensitive to the reaction CH3CH2CH2CH2OH + HO2 →
estimated based on the formalism of conventional transition
CH3CH2CH2C•HOH + H2O2. Likewise, it was found that
state theory (TST).41 To reflect the correct reaction
20% of the reaction fluxes during 1-butanol combustion pass
degeneracy, fitted pre-exponential A factors were multiplied
via the former reaction at 950 K and 5 atm. Reaction rate
by the number of equivalent abstractable hydrogen atoms.
constants for H abstraction from butanol isomers were Plausible effects of quantum tunneling on reaction rate
extrapolated from a theoretical prediction by Zhou et al.30 constants were accounted for with the inclusion of a one-
for the system 1-butanol + HO2. Other theoretically calculated dimensional Eckart functional as implemented in the KiSThelP
rate constants encompass values for the systems ethanol + code.42 The CH3- and OH- internal rotations in reactants and
HO231 and methanol + HO2.32 Reaction rate constants for the transition structures necessitate their rigorous treatment as
latter reaction are scattered with one order of magnitude. hindered rotors. The latter were obtained by performing partial
The significant importance of the H-abstraction reaction by optimizations along the corresponding dihedral angles. In
HO2 radicals in alcohol oxidation originates from the addition to CH3 and OH internal rotations in reactants and
formation of HO2 radicals and stable aldehyde species transition structures, we also treat the internal rotation of the
following O2 addition to the alpha fuel radical of alcohol.29,33 entire OH group in the HO2 moiety of the transition structures
The α-carbon site is defined as the one next to the OH moiety as hindered rotors. Treatment of hindered rotors in the
and is typically the weakest C/O−H site in alcohols. Due to KiSThelP package deploys the hindered rotor density-of-states
the appreciable amount of HO2 produced in this pathway, the (HRDS) approach formulated by McClurg et al.43 The only
H-abstraction reaction by HO2 is further promoted. To this required input parameter in the HRDS formalism is the ratio of
end, the aim of this study is to present thermal rate constants the overall rotational barrier to the associated harmonic
for H abstraction by HO2 radicals from various distinct sites in frequency.
C1−C5 primary alcohols. Kinetic parameters provided herein The Chemkin-Pro44 package predicts ignition delay times
shall be useful in improving model prediction of ignition delay using the commonly deployed constant-volume homogeneous
times, especially at low temperatures and high pressures. The batch reactor model. The ignition delay times correspond to
performance of literature kinetic models in predicting ignition the point of maximum temperature rise, max dT/dt, a point
delay times of primary alcohols is revisited in view of Arrhenius that also reflects the maximum pressure rate rise. Simulations
coefficients for HO2-based reactions computed herein. The are based on published chemical kinetic mechanisms and
11782 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

thermodynamic data for the investigated C1−5 alcohols with of literature kinetics models in predicting ignition delay times
the inclusion of modified kinetic parameters for HO2- of C1−5 primary alcohols based on modified HO2-abstraction
abstraction reactions. reactions computed herein. The phrase “this work” in all tables
and figures denotes updated kinetic parameters for HO2-
3. RESULTS AND DISCUSSIONS abstraction reactions only in corresponding literature kinetic
Figure 1 depicts optimized geometries for the 10 considered models. While the error bar on experimental data is not
primary C1−C5 alcohols along with bond dissociation displayed in figures that report ignition delay times herein, it is
generally accepted that ignition delay times in shock tubes and
rapid compression machines typically incur uncertainties of
around 2045 and 15%, in that order.46 It is worth mentioning
that all abstraction reactions considered herein from secondary
sites are associated with negative entropies of activation in the
range of ∼110 to 150 J/mol K between 500 and 2000 K. As an
illustrative example, Figure S1 in the Supporting Information
plots entropy of activation for H abstraction by HO2 from the
secondary site in ethanol. Consequently, entropies of activation
reduce the A factors in all investigated bimolecular reactions.
3.1. Methanol (CH3OH + HO2 → CH2OH + H2O2).
Screening analysis by Klippenstein et al.32 demonstrated that
the reaction CH3OH + HO2 → CH2OH + H2O2 largely
dominates uncertainty in the predicted ignition delay of
methanol based on the kinetic model of Li et al.47 The
noticeable effect of the kinetics of this reaction on the ignition
delay times of methanol flames is more profound at high
pressures and low temperatures, relevant to real combustion
engines (>10 atm and 1000−1200 K). The literature provides
no direct experimental measurement for this reaction. A very
slow anticipated rate for this reaction (at low temperatures)
and rapid decomposition of HO2 radicals (at high temper-
Figure 1. Structures of the four considered C1−3 alcohols. Values
atures) make experimental kinetics measurements very
signify bond dissociation enthalpies in kJ/mol.25 Considered challenging. As recently reviewed by Curran,9 the three
abstraction sites are highlighted. available theoretically derived rate constants18,32,40 for this
reaction significantly vary by almost an order of magnitude.
The noticeable difference in these estimates stems from the
enthalpies (BDHs) for C/O−H bonds as reported by Sarathy adapted theoretical frameworks pertinent to the calculation
et al.25 Herein, we compute reaction rate constants for H methodologies and treatment of anharmonicities in reactants
abstraction by HO2 radicals from the weakest sites. With the
and transition states. Our computed reaction and activation
noticeable difference in BDHs across the carbon skeleton of
energies for this reaction amount to 38 and 63 kJ/mol,
C1−5 primary alcohols, it is anticipated that enthalpic factors
respectively. These values compare very well with analogous
largely dominate entropic factors across the low to
values (34 and 64 kJ/mol) reported by Klippenstein et al.32 (at
intermediate temperatures intervals. These temperatures are
most pertinent to the auto-ignition window. To call attention the UCCSD(T)/CBS//CASPT2/cc-pvtz level of theory). The
for a plausible competition between different abstractable H calculated activation energy by Alecu and Truhlar40 (based on
sites, we highlight branching ratios for H abstraction from CCSD(T)/CBS formalisms) appears to exceed our computed
distinct sites in selected primary alcohols with comparable value in the range of 7−20 kJ/mol. Figure 3 contrasts our
BDHs values (i.e., within 15−20 kJ/mol). C−H bonds in the computed rate constant for the CH3OH + HO2 → CH2OH +
CH2 group attached (s/s1 sites) to the OH group in normal H2O2 reaction with other theoretically derived and simulated
alcohols are significantly weaker than other C−H sites. For values.9,18,32,40,48 In Figure 3, rate coefficients by Rasmussen et
instance, BDH for the s1 C−H site in n-propanol is lower by 20 al.48 came from fitting of H2O2 concentration profiles from
kJ/mol than the adjacent secondary C−H site, s2. BDH values methanol oxidation. At 1000 K, our computed rate constant
for tertiary C−H (t) in branched alcohols (i-butanol and i- exceeds analogous values by Klippenstein et al.32 and Alecu
pentanol) are very close to analogous values of s1 sites. Thus, and Truhlar40 by factors of 7.4 and 27.6, respectively. As
these two sites were considered. The portrayed geometries for pointed out by Curran,9 validation of the methanol oxidation
transition structures in Figure 2 share very similar features. For model against ignition delay times appears to agree better
instance, the O···O bond in the HO2 radical is consistently when deploying our 201118 computed reaction rate coefficient.
elongated by ∼8.0% in reference to the equilibrium distance in Our recalculated reaction rate constant for this reaction in this
HO2 radicals. Table 1 enlists fitted modified Arrhenius work remains close to our previously estimated value18 (a
parameters between 500 and 2000 K for H abstraction by difference by a factor of only 1.7). Such a difference partly
HO2 radicals from the highlighted sites in Figure 1. arises from a disparity of ∼15 kJ/mol in the estimated
In the following sections, we compare our calculated activation energy (63 kJ/mol) calculated herein at the CBS-
reaction rate constants with analogous literature values, either QB3 level of theory and our previously estimated value of 78
computed by DFT/ab initio methods or deployed in recent kJ/mol18 computed at the BB1K/GTLarge theoretical
kinetics models. The prime focus is to survey the performance approach.
11783 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Figure 2. Optimized structures for transition states. Distances are in Å.

Table 1. Computed Reaction (ΔH) and Activation Enthalpies (ΔH#) for H Abstraction by HO2 from C1−5 Primary Alcohols
and Their Arrhenius Parameters (Fitted in the Temperature Range 500−200 K)a
alcohol TS reaction ΔH ΔH# A n Ea
−30
methanol TS1 CH3OH + HO2 → CH2OH + H2O2 (3) 38 63 1.23 × 10 5.85 32.0
ethanol TS2 C2H5OH + HO2 → CH3CHOH + H2O2 (2) 32 51 1.04 × 10−33 6.63 15.2
n-propanol TS3 n-C3H7OH + HO2 → n(s1)-C3H6OH + H2O2 (2) 32 48 8.67 × 10−33 6.26 15.8
TS4 n-C3H7OH + HO2 → n(s2)-C3H6OH + H2O2 (2) 52 59 3.63 × 10−35 6.77 20.3
i-propanol TS5 i-C3H7OH + HO2 → i(t)-C3H6OH + H2O2 (1) 27 40 4.70 × 10−31 5.69 9.2
n-butanol TS6 n-C4H9OH + HO2 → n(s1)-C4H8OH + H2O2 (2) 32 49 2.07 × 10−27 4.65 24.0
2-butanol TS7 2-C4H9OH + HO2 → 2(s1)-C4H8OH + H2O2 (2) 38 63 1.79 × 10−27 4.71 37.3
TS8 2-C4H9OH + HO2 → 2(t)-C4H8OH + H2O2 (1) 29 51 3.34 × 10−27 4.17 16.6
i-butanol TS9 i-C5H12OH + HO2 → i(s)-C5H11OH + H2O2 (2) 34 44 7.71 × 10−27 4.20 22.4
TS10 i-C5H12OH + HO2 → i(t)-C5H11OH + H2O2 (1) 45 46 9.53 × 10−26 3.76 28.2
t-butanol TS11 C(CH3)3OH + HO2 → C(CH3)3O + H2O2 (9) 74 81 2.09 × 10−23 3.34 59.9
TS12 C(CH3)3OH + HO2 → C(CH3)2CH2OH + H2O2 (1) 64 66 8.24 × 10−25 3.74 49.5
i-pentanol TS13 i-C5H12OH + HO2 → i(s1)-C5H11OH + H2O2 (2) 30 47 2.26 × 10−29 5.09 19.6
TS14 i-C5H12OH + HO2 → i(s2)-C5H11OH + H2O2 (2) 59 74 9.74 × 10−29 5.02 45.5
TS15 i-C5H12OH + HO2 → i(t)-C5H11OH + H2O2 (1) 40 48 6.56 × 10−27 4.49 24.1
n-pentanol TS16 n-C5H12OH + HO2 → n(s1)-C5H11OH + H2O2 (2) 33 44 1.71 × 10−27 4.40 21.1
TS17 n-C5H12OH + HO2 → n(s3)-C5H11OH + H2O2 (2) 49 58 1.66 × 10−26 4.45 34.2
a
All values of energies and A are in the units kJ/mol and molecule cm3 s−1, respectively. Numbers in parenthesis denote reaction’s degeneracies.

The kinetic model by Burke et al.49 on methanol oxidation title reaction (CH3OH + HO2 → CH2OH + H2O2). The
utilized our previously computed kinetics parameters for the model reasonably predicted ignition delay times in reference to
11784 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Figure 3. Comparison between calculated reaction rate constants for


H abstraction from the methyl group in methanol by HO2 with
analogous literature values. aRef 18, bref 32, cref 40, dref 48, and eref 9.

experimental measurements in shock tubes by Noorani et al.50


and others. Between 800 and 1600 K, Our calculated rate
constants herein surpass values of Altarawneh et al.18 by 1.74−
3.56. Thus, we focus our attention now to comparing the
performance of the model of Burke et al.49 in estimating
ignition delay times of methanol with the inclusion of updated
and slightly faster kinetics coefficients for the reaction CH3OH
+ HO2 → CH2OH + H2O2. Figure 4 contrasts simulated

Figure 5. (a) Profiles of HO2 and OH from oxidation of methanol


and (b) relative % consumption of methanol by OH and HO2
radicals. Concentrations and relative % consumption are obtained at
fixed temperatures between 800 and 1200 K at 10 atm.

cantly exceed those of OH radicals up to a temperature as high


as 1100 K, attesting to the central role of HO2 in the initial
oxidation stages of methanol. Higher HO2 concentrations at
low temperatures was also quantified by Blocquet et al.13
Figure 4. Influence of the updated reaction rate constant for the
CH3OH + HO2 → CH2OH + H2O2 reaction on simulated ignition during oxidation of n-butane in Java specification requests
delay times (based on the model of Burke et al.49) for methanol in (JSR) based on the FAGE technique. Analysis of reaction
reference to analogous experimental data by Noorani et al. aRef 49 paths indicates prevalent contributions from the reactions
and bref 50. CH3OH + HO2 → CH2OH + H2O2 (∼800 K), HCO + O2 →
CO + HO2, CH2OH + O2 → CH2O + HO2 (∼850 K), and H
+ O2(+M) → HO2(+M) (>900 K) toward formation of HO2.
ignition delay times of stoichiometric flames of methanol at a The relative consumption of methanol by OH and HO2
dilution of D = 20 (in Ar) against experimental measurements radicals is portrayed in Figure 5b between 800 and 1200 K
in a shock tube by Noorani et al.50 At 1068 and 1098 K, the at 10 atm. At 900 and 1000 K, HO2 radicals account for 17 and
original model of Burke et al.49 overpredicts experimental 23% of the initial consumption of methanol. Following H
measurements by 58 and 27%, respectively. Considering the abstraction from methanol by oxygen molecules, the low-
updated kinetics for the title reaction calculated herein, temperature oxidative decomposition of methanol (i.e., 800−
simulated values by the model of Burke et al.49 deviate by 850 K) is chiefly derived from HO2 radicals but approaches
only 1 and 19% from the analogous experimental measure- zero at ∼1220 K. Across the low-temperature window, the
ments. The original model by Burke et al.49 affords a better main exit channel for HO2 radicals ensues via their self-
comparison with experimental values at intermediate temper- combination along the reaction 2HO2 → O2 + H2O2.
atures. Nonetheless, the role of HO2 in deriving auto-ignition Hydrogen peroxide radicals decompose into OH radicals,
significantly diminishes as the temperature increases. thus initiating the auto-ignition. It is inferred from Figure 5
To attain further insight into the role of HO2 in controlling that auto-ignition of methanol takes place via the gradual
ignition delay times of methanol, we plot in Figure 5a modeled buildup of HO2 radicals and their subsequent decomposition
concentration profiles of OH and HO2 radicals from oxidation into OH radicals. The role of HO2 radicals in combustion of
of methanol. The predicted concentrations of HO2 signifi- methanol effectively diminishes after the onset of the auto-
11785 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

ignition. The P/T-dependent contribution of HO2 radicals in


the initial oxidation of alcohols deserves further careful
scrutiny.
3.2. Ethanol (C2H5OH + HO2 → s-C2H4OH + H2O2).
Among C1−5 primary alcohols, combustion of ethanol is
arguably the most investigated system. Combustion character-
istics of ethanol (flame speeds, ignition delay times, and
profiles of species) were validated in numerous kinetic
models.51−54 A noticeable discrepancy among these models
in predicting ignition delay times of ethanol was often
attributed to considerable uncertainty in the kinetics of HO2
reactions, especially at low temperatures and high pressures.
Predicted ignition delay times in these models show a rather
moderate agreement with corresponding experimental meas-
urements. Kinetic parameters for the reaction C2H5OH + HO2 Figure 6. Comparison between calculated reaction rate constants for
→ s-C2H4OH + H2O2 in these models are often extrapolated H abstraction from a secondary C−H site in ethanol by HO2 radicals
by analogy to corresponding values computed for the n- with respective literature values. aRef 30, bref 31, cref 54, and dref 52.
butanol + HO2 system by Zhou et al.30 Very recently, Zhao et
al.31 deployed the CCSD(T)/CBS//M062X/def-TZVP the- order of magnitude. With a difference of only 5 kJ/mol in the
oretical approach to compute rate constants for H abstractions calculated ΔH# value (translates to a difference of only a factor
by HO2 radicals from the three distinct sites in ethanol. It was of 0.5 in k (1000 K) values), the noticeable difference in the
found that H abstraction from the ethanol proceeds at slower reaction rate constant must stem from the treatment of
rates when compared with analogous values for the n-butanol + anharmonicity. As stated in section 2, we have treated all
HO2 system. For instance, H abstraction by HO2 from the internal CH3/OH rotations in reactants and transition states
secondary site in n-butanol exceeds analogous values in ethanol (including the internal rotation of the H−O bond in transition
by factors within ∼37 to 23 between 800 and 1200 K. Zhao et states) as hindered rotors. For instance, Figure S2 in the
al.31 demonstrated that updating kinetics for ethanol + HO2 Supporting Information displays potential energy diagrams for
reactions displayed a rather limited effect on ethanol reactivity internal rotations of CH3 and OH groups in selected transition
at low temperatures and high pressures (in reference to models structures. It is not clear from the description of methodology
that utilize values calculated for the n-butanol + HO2 system). by Zhao et al.31 if torsional modes in transition states were also
A further improvement could be potentially attained by treated as hindered rotors. It is inferred from Figure 6 that
improving the acetaldehyde and vinyl alcohol submechanisms computed rate constants are higher than all other literature
and by reassessing kinetics for the other important initiation values. Nonetheless, other theoretically obtained values by
C2H5OH + O2 reactions. Zhao et al.31 (for the ethanol + HO2 system) also differ from
Among the various kinetic models on ethanol oxidation, the analogous values by Zhou et al.30 (for the n-butanol + HO2
detailed combustion chemistry model by Mittal et al.52 system) by a factor of ∼2 between 800 and 1200 K.
reasonably reproduced salient combustion features of ethanol. We next turn our attention to testing the predictive capacity
The model by Mittal et al.52 is based on the C1−C2 mechanism of the model by Mittal et al.52 with updated (and a markedly
by Metcalfe et al.,55 (AramcoMech 1.3.). To improve the faster) kinetic parameters for the C2H5OH + HO2 → s-
performance of the model in predicting ignition delay times, C2H4OH + H2O2 reaction. In Figure 7a, predicted ignition
the A factor for the reaction C2H5OH + HO2 → s-C2H4OH + delay times by our updated model of Mittal et al.52
H2O2 (extrapolated from the n-butanol + HO2 system) was underestimate analogous experimental values in comparison
increased by a factor of 1.75. Barraza-Botet et al.54 utilized the to the original models of Mittal et al.52 and Barraza-Botet et
model of Mittal et al.52 to account for measured ignition delay al.54 Values by the updated model of Mittal et al.52 approach
times of ethanol in a rapid compression facility. Likewise, in estimates of the updated model beyond 1000 K. On the
the model by Barraza-Botet et al.,54 the A factor for the title contrary and as Figure 7b shows, the updated model of Mittal
reaction was enhanced by a factor of 2.5 in order to improve et al.52 affords an improved match with experimental values by
the predictive performance of the model against experimental Noorani et al.50 in reference to the original models of Mittal et
measurements of the ignition delay times. The 1.75/2.5 factors al.52 and Barraza-Botet et al.54 In both cases (Figure 7a,b),
were within the uncertainty limits of the calculations by Zhou utilizing the recent calculated reaction constant parameters by
et al.30 on the n-butanol + HO2 system. Figure 6 compares our Zhao et al.31 results in a significant overprediction of the
computed Arrhenius plots for this reaction against correspond- ignition delay times in comparison to the three other models.
ing values used in the models of Mittal et al.52 and Barraza- Figure 8 shows results of sensitivity analysis on formation of
Botet et al.54 and calculated values by Zhou et al.30 and Zhao OH and HO2 radicals during ignition of a stoichiometric
et al.31 ethanol mixture at 1 atm under air dilution (D = 3.76). The
Our computed values depart from those of Zhou et al.30 at prominent role of HO2 radicals in controlling the ignition
the most relevant temperature to auto-ignition (800−1200 K) delay times is clearly illustrated in Figure 8a,b. The most
by factors of 7.9−5.2. Consequently, our calculated rate important reactions for the consumption of HO2 radicals are
constants for this reaction exceed deployed values in the also the most prominent reactions in the formation of OH
models of Barraza-Botet et al.54 and Mittal et al.52 by 3.2−2.1 radicals. HO2 radicals are mainly consumed through the two
and 4.5−3.0, respectively (between 800 and 1200 K). As reactions C2H5OH + HO2 → s-C2H4OH + H2O2 and
Figure 6 portrays, our calculated k(T) values significantly H2O2(+M) → 2OH(+M). These two reactions are also the
surpass corresponding values by Zhao et al.31 by almost an two most significant reactions toward production of OH
11786 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Figure 7. Influence of updated reaction rate constants for the


C2H5OH + HO2 → s-C2H4OH + H2O2 reaction on simulated
ignition delay times (based on the model of Mittal et al.52) in
reference to analogous experimental data by (a) Barazza-Botet et al.54
and (b) Noorani et al.50 aRef 54, bref 52, cref 30, and dref 50.

radicals. As Figure 8c depicts, the sharp decrease in the


concentration of HO2 radicals at the time of ignition
accompanies the sudden increase in the concentration of
OH radicals.
3.3. Isomers of Propanol (n-C3H7OH/i-C3H7OH). When
compared with methanol and ethanol, isomers of propanol
attain a higher energy density and reduced evaporative
emission profiles. As the simplest “next-generation” alcohols,
they mimic to a large extent the combustion property of real
biofuels. Several shock tube experiments reported ignition
delay times for the isomers of propanol.50,56−58 With a shorter
ignition delay time, n-propanol is more reactive than i-
propanol.25 This unique observation was mainly attributed to a
relatively weaker secondary C−H bond in the former and to Figure 8. Sensitivity coefficients for the 10 most important reactions
the formation of highly reactive formaldehyde in the flame of i- leading to formation/consumption of (a) OH and (b) HO2 during
propanol as opposed to the less reactive acetone in the case of combustion of ethanol. (c) Plots of the concentration of HO2 and
i-propanol. The literature presents no theoretical estimates for OH around the ignition window.
H abstraction from propanol isomers by HO2 radicals. In
simulating ignition delay times, kinetic models on combustion H abstraction from the s1 site in n-propanol and H abstraction
of i-propanol and n-propanol utilized reaction rate constants from the secondary site in ethanol ensue via very comparable
for HO2-based reactions from analogous values calculated for rates at all investigated temperatures.
n-butane and alkanes. The model of Man et al.56 updated the Figure 9 compares our computed rate constant for H
C3 chemistry in the study of Johnson et al.57 with a prime focus abstraction from secondary C−H and tertiary C−H sites in n-
on unimolecular initiation reactions and H abstraction propanol and i-propanol, respectively, with their analogous
reactions by the O/H radical pool. H abstraction from the t literature values.17,56 Between 800 and 1200 K, calculated rate
site in i-propanol exceeds that of H abstraction from the s1 site constants herein for n-propanol are within factors of ∼2.0 to
in n-propanol by factors within 8−2 between 500 and 1200 K. 1.17 from values deployed in the model of Man et al.56
Owing to similar activation enthalpies (51 versus 48 kJ/mol), Likewise, our computed values for H abstraction from i-
11787 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

propanol exceed values utilized in the model of Man et al56 by


factors of 11−4.6 between 800 and 1200 K. As illustrated in
the previous section on methanol, the difference in calculated
rate constants partially originates from the methodology of
treatment of anharmonicities. The noticeable difference with
values computed by Aguilera-Iparraguirre et al. 17 for
corresponding alkanes (n-C3H10 and i-C4H10) originates from
distinct activation energies. For instance, computed activation
energy for H abstraction from the tertiary site in i-butane
exceeds analogous values in i-propanol by 14 kJ/mol. Similarly,
activation energy for H abstraction from the s1 site in n-
propanol is lower by 16 kJ/mol if compared with H abstraction
from the secondary site in propane. Our calculated activation
energy for H abstraction from the s2 site in n-propanol
perfectly matches the corresponding value in the case of
propane (59 versus 60 kJ/mol). As such, calculated reaction
rate constants for H abstraction from the s2 site in n-propanol
remain within a factor of only ∼2 from analogous values
computed by Aguilera-Iparraguirre et al.17 for H abstraction
from the secondary site in n-propane.
In Tables 2 and 3, we compare the predictive performance
of the original model of Man et al.56 and the updated model of
Man et al.56 (in which the kinetics of HO2-based abstraction
reactions are modified based on current calculations) against
experimental measurements from different groups50,56,58 at
various combinations of temperatures and pressures. As shown
in Tables 2 (n-propanol) and 3 (i-propanol), the updated
model of Man et al.56 only slightly improves the overprediction
of ignition delay times in reference to the original model. At
temperatures above 1200 K, both models essentially yield very
similar values. The very similar performance of original and
Figure 9. Comparison between calculated reaction rate constants for
H abstraction from the secondary and tertiary sites in (a) n-propanol updated models of Man et al.56 is clearly illustrated in Figure
and (b) i-propanol by HO2 radicals with corresponding literature 10a,b with both models overpredicting experimentally fitted
values. aRef 56 and bref 17. values by Noorani et al.50 Nonetheless, both models still
outperform the model of Johnson et al.57 The updated model
correctly reproduces the confirmed lower reactivity of i-

Table 2. Comparison between Experimental and Simulated Values of Ignition Delay Times for n-Propanola
P (atm) T (K) Jouzdani et al.58 (exptl) Man et al.56 (exptl) Noorani et al.50 (exptl) this work Man et al.56 (model)
3.5 1152 2017 (122%) 4482 4590
2.4 1155 1724 (250%) 5597 5770
3.6 1163 2097 (80%) 3767 3840
1.28 1185 2181 (141%) 5370 5435
3.7 1206 1379 (47%) 2020 2050
3.8 1208 1271 (52%) 1940 1974
3.7 1210 1179 (39%) 1920 1944
3.7 1227 1005 (52%) 1520 1530
1.2 1230 1251 (130%) 2880 2898
2.3 1236 889 (50%) 1790 1800
3.5 1267 687 (33%) 914 918
1.17 1288 656 (140%) 1410 1320
3.3 1313 400 (31%) 522 523
2.2 1341 420 (12%) 470 472
3.3 1372 241 (7%) 258 262
2.1 1413 336 (37%) 212 174
1.25 1423 196 (34%) 262 262
3.3 1458 82 (26%) 103 104
3.1 1490 66 (19%) 78 78
1.28 1554 67 (18%) 79 77

a
ϕ = 1.0, D = 21(in Ar). Percentage values denote differences between experimental values and values calculated with our updated model (based
on the model of Man et al.56).

11788 DOI: 10.1021/acs.energyfuels.9b02169


Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Table 3. Comparison between Experimental and Simulated


Values of Ignition Delay Times for n-Propanola
Jouzdani Man et al.56 this Man et al.56
P (atm) T (K) et al.58 (exptl) (exptl) work model
17.29 1192 1014 (182%) 2910 3507
3.4 1240 2203 (65%) 3640 3938
3.1 1297 1206 2020 2097
4.13 1300 931 (71%) 1593 1661
3.5 1329 851 (48%) 1260 1297
3.9 1419 303 (32%) 399 403
17.52 1450 72 (67%) 120 123
3.2 1468 255 (1%) 254 254
3.6 1553 73 (33%) 97 98
1.23 1598 86 (41%) 121 119
a
ϕ = 1.0, D = 21 (in Ar). Percentage values denote differences
between experimental values and values calculated with the updated
model (based on the model of Man et al.56).

propanol in reference to n-propanol (Figure 10c). Matching


the experimentally reported ignition delay times for both
isomers of propanol necessitates significantly faster HO2-
abstraction reactions and improving the C3 submechanism.
3.4. Isomers of Butanol (n-C4H9OH/2-C4H 9OH/i-
C4H9OH/t-C4H9OH). The seminal paper by Sarathy et al.29
presented a detail combustion chemistry model for the four
isomers of butanol. The model satisfactorily accounts for the
species profile, ignition delay times, and laminar flame speeds
under a wide range of operational conditions (temperatures,
pressures, dilution extent, and equivalence ratios) encountered
in various combustion devices (rapid compression machines,
shock tubes, JSR, and low-pressure premixed flames). The
model is based on the hierarchical nature of the combustion
chemistry starting from the H2/CO/O2 submechanisms while
comprehensively updating and improving the C1−C3 mecha-
nisms. The model distinguishes between reactions prevailing in
the low-temperature window from that dominating in high-
temperature oxidation. As expected, ignition delay times were
the most sensitive to HO2-abstraction reactions based on brute
force sensitivity analysis. HO2 radicals accounted for 20%
consumption of n-butanol at 950 K and 5 atm under
stoichiometric conditions. In the model of Sarathy et al.,29
rate constants for n-butanol + HO2 reactions were adapted
from the study of Zhou et al.30 (with the A factor multiplied by
2.5). For the other butanol isomers, reaction rate constants
were extrapolated from the n-butanol + HO2 system based on Figure 10. Ignition delay times for (a) n-propanol and (b) i-propanol
an Evans−Polanyi-type correlation. For the t-butanol molecule and (c) comparison between them. Our calculated values are based
on the model of Man et al.56 with a modified rate constant for H
in particular, the Evans−Polanyi correlation yields a reaction abstraction by HO2. aRef 57, bref 56, and cref 50.
rate constant that resulted in very long ignition delay times.
Thus, rate constant for H abstraction from the primary site in
t-butanol by HO2 radicals was assigned with the same A factor
computed in the case of n-butanol. times of n-butanol when compared with experimental
In Figure 11, we compare our computed reaction rate measurements by Noorani et al.50 between 1150 and 1500 K
constants for H abstraction from the weakest C−H sites in the and 8.8−11.7 atm. However, at low to intermediate temper-
four butanol isomers with corresponding values deployed in atures (∼800 to 1000 K), our calculated rate constants for H
the model of Sarathy et al.29 For abstraction from the s1 site in abstraction from the s1 site in n-butanol exceed values deployed
n-butanol, the two sets of values fall within factors of 1.3−1.0 in the model of Sarathy et al.29 by factors between ∼3 and 2.
between 1200 and 1500 K. As such, updating the reaction rate This noticeable difference causes the updated model of Sarathy
constant for the reaction n-C4H9OH + HO2 → n(s1)-C4H8OH et al.29 to perform better at low temperature in predicting
+ H2O2 in the model of Sarathy et al.29 derives a minimal ignition delay times. In Figure 12b, the original model by
difference in predicted ignition delay times at temperatures Sarathy et al.29 significantly over predicts ignition delay times
above ∼1150 K. As Figure 12a shows, the original and updated measured by Zhu et al.59 at low temperatures and high
model by Sarathy et al.29 slightly overpredicts ignition delay pressures. Simulated values based on the modified model by
11789 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Figure 11. Comparison between calculated reaction rate constants in this work with analogous literature values for (a) n-butanol, (b) 2-butanol,
(c) i-butanol, and (d) tetra-butanol. aRef 30 and bref 29.

Sarathy et al.29 reside within 20−50% of the corresponding experimental measurements by Weber and Sung.60 Nonethe-
experimental values. less, the original model by Sarathy et al.29 retains a better
Inspection of values in Figure 11b for 2-butanol reveals a predictive performance for i-butanol as Figure 13b shows.
fundamental difference between our computed reaction rate With these striking differences in absolute values of reaction
constants herein and values deployed in the model of Sarathy rate constants and relative importance for specific abstractable
et al.29 Between 800 and 1200 K, the contribution of H sites, it is insightful to compare the overall predictive
abstraction from the tertiary site decreases from 0.53 to 0.24. performance of the original model of Sarathy et al.29 with an
In the model of Sarathy et al.,29 abstraction from the tertiary updated model that includes modified rate constants for HO2-
site largely dominates that of the secondary site with branching abstraction reactions. Ignition delay times for the four isomers
ratios in the range of 0.96 (800 K) to 0.87 (1200 K). For the i- depicted in Figure 14 (based on the updated HO2 rate
butanol isomer, the tertiary C−H bond is stronger than the constants) reflect to a large extent the trend deduced from the
secondary C−H bond by 9 kJ/mol. Thus, it is expected that original model of Sarathy et al.29 Above 850 K and at 15 atm
abstraction from the secondary site proceeds at a faster rate. (φ = 1.0 with air dilution), the original model of Sarathy et
Indeed, this is the trend that Figure 11c illustrates. The al.29 predicts n-butanol to be the most reactive isomer with the
branching ratio for abstraction from the secondary site remains shortest ignition delay times. The t-butanol isomer exhibits the
consistently high at ∼0.78 (800−1200 K). On the contrary, least reactivity, while 2-butanol and i-butanol isomers incur
abstraction from the tertiary site dominates the i-butanol + very comparable reactivity.
HO2 reactions in the model of Sarathy et al.29 with a branching 3.5. Isomers of Pentanol (n-C5H11OH/i-C5H11OH). With
ratio at 0.97 (800−1200 K). Finally, our calculated rate similar physiochemical properties to gasoline, i-pentanol (3-
constants for abstraction from the primary site in t-butanol are methyl-1-butanol) has emerged as a potential bioderived fuel.
approximately one order of magnitude slower than analogous Tsujimura et al.61 investigated the auto-ignition character of i-
values utilized in the model of Sarathy et al.29 However, values pentanol in shock tube and rapid compression machine
deployed in the model of Sarathy et al.29 exceed corresponding experiments under a wide range of equivalence ratios,
values for H abstraction from primary sites in alkanes by a temperatures, and pressures. The deployed kinetic model is
factor of 4 at 1000 K. Calculated reaction constants herein largely based on existing models for C1−C4 alcohols with the
improve the predictive performance for the model of Sarathy et inclusion of updated reactions rate constants for a concerted
al.29 for t-butanol at low temperatures (800−860 K) and high HO2 elimination step from ROO-type radicals along the
pressures (15 atm) as Figure 13a portrays, in reference to reaction scheme proposed by da Silva et al.62 for ethanol. The
11790 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

Figure 12. Comparison between computed ignition delay times for n-


butanol with corresponding experimental values at (a) high and (b)
low temperatures. aRef 50, bref 29, and cref 59. Computed ignition
delay times herein are based on the model of Sarathy et al.29 with
modified rate constants for HO2 reactions with the four isomers. Figure 13. Influence of updated reaction rate constants for H
abstraction from (a) t-butanol and (b) i-butanol on simulated ignition
model satisfactorily describes the monotonous decrease in delay times (based on the model of Sarathy et al.29) in reference to
respective experimental data by Weber and Sung.60 aRef 60 and bref
ignition delay times with temperature. Rate coefficients for
29.
abstraction reactions by HO2 radicals were adapted from
analogous values calculated by Aguilera-Iparraguirre et al.17 for
alkanes. To improve comparison with experimental measure-
ments of ignition delay times, the A factor in these reactions
was multiplied by a factor of 3.0. In another study, Tang et al.63
investigated high-temperature ignition delay times for C5
primary alcohols (i-pentanol, n-pentanol, and 2-methyl-1-
butanol). Ignition delay times were simulated based on kinetic
models developed by Dagaut and co-authors.64,65
Figure 15 contrasts our computed rate constants for H
abstraction by HO2 radicals from distinct sites in i-pentanol
and n-pentanol with analogous values deployed in the model of
Tsujimura et al.61 Our calculations indicate that abstraction
from the tertiary site prevails over abstraction from the s1 site
in i-pentanol at all temperatures. The branching ratio for
abstraction from the s1 site remains constant at ∼0.26 between Figure 14. Ignition delay times for the four butanol isomers.
800 and 1200 K. In the model of Tsujimura et al.,61 abstraction
from the s1 sites dominates the overall i-pentanol + HO2
reactions at a branching ratio of ∼0.70 between 800 and 1200 models satisfactorily reproduce experimentally measured
K. Despite these differences, our computed rate constants for ignition delay times at 7 and 20 atm between 800 and 1600 K.
the overall reaction i-pentanol + HO2 reside within factors of Overall, we envision that the difference between simulated
2.5−1.3 between 800 and 12 00 K from corresponding values and experimental values is underpinned by the anticipated
deployed in the model of Tsujimura et al.61 Consequently, the uncertainties in experimental values of ignition delay times
updated model of Tsujimura et al.61 yields very similar ignition (typically within 15−20%) and by the expected aced accuracy
delay times to the original model. As Figure 16 shows, both margin of reaction rate constants calculated herein (i.e., a
11791 DOI: 10.1021/acs.energyfuels.9b02169
Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

decomposition of methanol between 900 and 1000 K at 10


atm. Slightly higher rate constants for the CH3OH + HO2
reaction improves prediction of ignition delay of methanol at
low temperatures and high pressures. Our calculated rate
constants for the reaction C2H5OH + HO2 → s-C2H4OH +
H2O2 significantly exceed very recent theoretical calculations.
Nonetheless, predicted ignition delay times of ethanol based
on our updated kinetic parameters for the former reaction
afford an improved match with analogous experimental
measurements, especially at high pressures and high dilution.
The role of HO2 in controlling ignition delay times of ethanol
manifests itself by the fact that the most important reactions in
depletion of HO2 radicals are also the most prominent reaction
in the formation of OH radicals. Calculated rate constants for
the reactions of HO2 with n-propanol/i-propanol slightly
Figure 15. Comparison between calculated Arrhenius plots for the improve the overprediction of their ignition delay times in
abstraction channels from i-pentanol and their corresponding values reference to the original literature kinetic model. While
used in the model of Tsujimura et al.61 aRef 61. s1, s2, and t signify computed rate constants for reactions of the butane isomers
abstraction from these distinct sites (shown in Figure 1 for i- with HO2 differ from deployed values in literature kinetic
pentanol). Solid lines refer to values calculated herein, while dashed models in absolute values and the nature of preferred
lines refer to values deployed in the model of Tsujimura et al.61 abstraction sites, our updated model accurately reproduces
ordering of the butane isomers reactivity. For i-pentanol, we
found that abstraction from the tertiary site prevails over that
of the secondary site. Ignition delay times of i-pentanol (based
on modified rate constants for HO2-abstraction reactions
herein) are very similar to values computed in the original
literature model. It is interesting from the perspectives of
modelers to compare the rate constants of various alcohols and
to compare the impact of chain length and branching
structures on these values. Such interesting aspects will be
addressed in due course.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
Figure 16. Comparison between computed ignition delay times for i- ACS Publications website at DOI: 10.1021/acs.energy-
pentanol with matching literature values. Computed values are based fuels.9b02169.
on the model by Tsujimura et al.61 with modified reaction rate
constants for HO2-abstraction reactions. aRef 61. Cartesian coordinates for transition structures (in Å)
(PDF)
factor of 3−5). A systematic benchmarking of ignition delay
times against HO2-derived rate constants by other chemistry
models (most notably, G4 and CBS-APNO) may improve the
■ AUTHOR INFORMATION
Corresponding Author
agreement between computed and measured ignition delay *E-mail: mn.altarawneh@uaeu.ac.ae.
times. Nonetheless, the ignition delay times in the investigated ORCID
alcohols might be sensitive to other sets of reactions. For
instance, a further improvement could be potentially attained Mansour H. Almatarneh: 0000-0002-2863-6487
by improving the acetaldehyde and vinyl alcohol submechan- Mohammednoor Altarawneh: 0000-0002-2832-3886
isms and by reassessing the kinetics for the other important Notes
initiation alcohols + O2 reactions. The authors declare no competing financial interest.

4. CONCLUSIONS
Reaction rate parameters are provided herein for H abstraction
■ ACKNOWLEDGMENTS
This work was supported by the Australian Research Council
from the weakest sites in C1−5 primary alcohols. Computed (ARC). We acknowledge the Pawsey Supercomputing Centre
activation enthalpies positively correlate with values of bond in Perth as well as the National Computational Infrastructure
dissociation enthalpies for the respective abstractable C−H (NCI) in Canberra, Australia, for providing the grants of
sites. For instance, a difference in bond dissociation enthalpies computational resources.
for the C−H bond in methanol and C−H in ethanol (the s
site) of 16 kJ/mol translates into a difference of 12 kJ/mol in
computed enthalpies of activation. The computational
■ REFERENCES
(1) Battin-Leclerc, F.; Herbinet, O.; Glaude, P.-A.; Fournet, R.;
methodology comprises CBS-QB3 energy estimation, hindered Zhou, Z.; Deng, L.; Guo, H.; Xie, M.; Qi, F. Experimental
rotor treatment, and a one-dimensional Eckart functional. We Confirmation of the Low-Temperature Oxidation Scheme of Alkanes.
found that HO2 radicals account for 17−23% of the initial Angew. Chem., Int. Ed. Engl. 2010, 49, 3169−3172.

11792 DOI: 10.1021/acs.energyfuels.9b02169


Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

(2) Battin-Leclerc, F. Detailed chemical kinetic models for the low- (20) Batiha, M.; Altarawneh, M.; Al-Harahsheh, M.; Altarawneh, I.;
temperature combustion of hydrocarbons with application to gasoline Rawadieh, S. Theoretical derivation for reaction rate constants of H
and diesel fuel surrogates. Prog. Energy Combust. Sci. 2008, 34, 440− abstraction from thiophenol by the H/O radical pool. Comput. Theor.
498. Chem. 2011, 970, 1−5.
(3) Zhang, P.; Ji, W.; He, T.; He, X.; Wang, Z.; Yang, B.; Law, C. K. (21) Altarawneh, M. K.; Dlugogorski, B. Z.; Kennedy, E. M.; Mackie,
First-stage ignition delay in the negative temperature coefficient J. C. Rate constants for reactions of ethylbenzene with hydroperoxyl
behavior: Experiment and simulation. Combust. Flame 2016, 167, 14− radical. Combust. Flame 2013, 160, 9−16.
23. (22) Altarawneh, M.; Dlugogorski, B. Z. Reactions of HO2 with n-
(4) Kurimoto, N.; Brumfield, B.; Yang, X.; Wada, T.; Diévart, P.; propylbenzene and its phenylpropyl radicals. Combust. Flame 2015,
Wysocki, G.; Ju, Y. Quantitative measurements of HO2/H2O2 and 162, 1406−1416.
intermediate species in low and intermediate temperature oxidation of (23) Zhang, J.; Niu, S.; Zhang, Y.; Tang, C.; Jiang, X.; Hu, E.;
dimethyl ether. Proc. Combust. Inst. 2015, 35, 457−464. Huang, Z. Experimental and modeling study of the auto-ignition of n-
(5) Ranzi, E.; Cavallotti, C.; Cuoci, A.; Frassoldati, A.; Pelucchi, M.; heptane/n-butanol mixtures. Combust. Flame 2013, 160, 31−39.
Faravelli, T. New reaction classes in the kinetic modeling of low (24) Gu, X.; Huang, Z.; Wu, S.; Li, Q. Laminar burning velocities
temperature oxidation of n-alkanes. Combust. Flame 2015, 162, 1679− and flame instabilities of butanol isomers−air mixtures. Combust.
1691. Flame 2010, 157, 2318−2325.
(6) Zádor, J.; Taatjes, C. A.; Fernandes, R. X. Kinetics of elementary (25) Sarathy, S. M.; Oßwald, P.; Hansen, N.; Kohse-Höinghaus, K.
reactions in low-temperature autoignition chemistry. Prog. Energy Alcohol combustion chemistry. Prog. Energy Combust. Sci. 2014, 44,
Combust. Sci. 2011, 37, 371−421. 40−102.
(7) Merchant, S. S.; Goldsmith, C. F.; Vandeputte, A. G.; Burke, M. (26) Pan, L.; Zhang, Y.; Tian, Z.; Yang, F.; Huang, Z. Experimental
P.; Klippenstein, S. J.; Green, W. H. Understanding low-temperature and Kinetic Study on Ignition Delay Times of iso-Butanol. Energy
first-stage ignition delay: Propane. Combust. Flame 2015, 162, 3658− Fuels 2014, 28, 2160−2169.
3673. (27) Yang, K.; Zhan, C.; Man, X.; Guan, L.; Huang, Z.; Tang, C.
(8) Mao, Y.; Yu, L.; Wu, Z.; Tao, W.; Wang, S.; Ruan, C.; Zhu, L.; Shock Tube Study on Propanal Ignition and the Comparison to
Lu, X. Experimental and kinetic modeling study of ignition Propane, n-Propanol, and i-Propanol. Energy Fuels 2016, 30, 717−
characteristics of RP-3 kerosene over low-to-high temperature ranges 724.
in a heated rapid compression machine and a heated shock tube. (28) Olm, C.; Varga, T.; Valkó, É .; Curran, H. J.; Turányi, T.
Combust. Flame 2019, 203, 157−169. Uncertainty quantification of a newly optimized methanol and
(9) Curran, H. J. Developing detailed chemical kinetic mechanisms formaldehyde combustion mechanism. Combust. Flame 2017, 186,
for fuel combustion. Proc. Combust. Inst. 2019, 37, 57−81. 45−64.
(10) Bahrini, C.; Herbinet, O.; Glaude, P.-A.; Schoemaecker, C.; (29) Sarathy, S. M.; Vranckx, S.; Yasunaga, K.; Mehl, M.; Oßwald,
Fittschen, C.; Battin-Leclerc, F. Quantification of Hydrogen Peroxide P.; Metcalfe, W. K.; Westbrook, C. K.; Pitz, W. J.; Kohse-Höinghaus,
during the Low-Temperature Oxidation of Alkanes. J. Am. Chem. Soc. K.; Fernandes, R. X.; Curran, H. J. A comprehensive chemical kinetic
2012, 134, 11944−11947. combustion model for the four butanol isomers. Combust. Flame
(11) Onel, L.; Brennan, A.; Gianella, M.; Ronnie, G.; Lawry Aguila, 2012, 159, 2028−2055.
A.; Hancock, G.; Whalley, L.; Seakins, P. W.; Ritchie, G. A. D.; Heard, (30) Zhou, C.-W.; Simmie, J. M.; Curran, H. J. Rate constants for
D. E. An intercomparison of HO2 measurements by fluorescence hydrogen abstraction by HȮ 2 from n-butanol. Int. J. Chem. Kinet.
assay by gas expansion and cavity ring-down spectroscopy within 2012, 44, 155−164.
HIRAC (Highly Instrumented Reactor for Atmospheric Chemistry). (31) Zhao, Q.; Zhang, Y.; Sun, W.; Deng, F.; Yang, F.; Huang, Z.
Atmos. Meas. Tech. 2017, 10, 4877−4894. Chemical Kinetics of H-Atom Abstraction from Ethanol by HȮ 2:
(12) Brumfield, B.; Sun, W.; Ju, Y.; Wysocki, G. Direct In Situ Implication for Combustion Modeling. J. Phys. Chem. A. 2019, 123,
Quantification of HO2 from a Flow Reactor. J. Phys. Chem. Lett. 2013, 971−982.
4, 872−876. (32) Klippenstein, S. J.; Harding, L. B.; Davis, M. J.; Tomlin, A. S.;
(13) Blocquet, M.; Schoemaecker, C.; Amedro, D.; Herbinet, O.; Skodje, R. T. Uncertainty driven theoretical kinetics studies for
Battin-Leclerc, F.; Fittschen, C. Quantification of OH and HO2 CH3OH ignition: HO2+CH3OH and O2+CH3OH. Proc. Combust.
radicals during the low-temperature oxidation of hydrocarbons by Inst. 2011, 33, 351−357.
Fluorescence Assay by Gas Expansion technique. Proc. Natl. Acad. Sci. (33) Cai, L.; Uygun, Y.; Togbé, C.; Pitsch, H.; Olivier, H.; Dagaut,
U. S. A. 2013, 110, 20014−20017. P.; Sarathy, S. M. An experimental and modeling study of n-octanol
(14) Jemi-Alade, A. A.; Lightfoot, P. D.; Lesclaux, R. A flash combustion. Proc. Combust. Inst. 2015, 35, 419−427.
photolysis study of the HO2 + HCHO → H2O2 + HCO reaction (34) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
between 541 and 656 K. Chem. Phys. Lett. 1992, 195, 25−30. Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
(15) Walker, R. W. Reactions of HO2 radicals in combustion B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
chemistry. Symp. Int. Combust. Proc. 1989, 22, 883−892. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
(16) Handford-Styring, S. M.; Walker, R. W. Rate constants for the Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.;
reaction of HO2 radicals with cyclopentane and propane between 673 Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;
and 783 K. Phys. Chem. Chem. Phys. 2002, 4, 620−627. Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.;
(17) Aguilera-Iparraguirre, J.; Curran, H. J.; Klopper, W.; Simmie, J. Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi,
M. Accurate Benchmark Calculation of the Reaction Barrier Height R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar,
for Hydrogen Abstraction by the Hydroperoxyl Radical from S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox,
Methane. Implications for CnH2n+2wheren= 2 → 4. J. Phys. Chem. J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.;
A 2008, 112, 7047. Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.;
(18) Altarawneh, M.; al-Muhtaseb, A. H.; Dlugogorski, B. Z.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.;
Kennedy, E. M.; Mackie, J. C. Rate constants for hydrogen abstraction Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A.
reactions by the hydroperoxyl radical from methanol, ethenol, D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.
acetaldehyde, toluene, and phenol. J. Comput. Chem. 2011, 32, Gaussian 09, Revision A.02; Gaussian, Inc.: Wallingford CT, 2009.
1725−1733. (35) Montgomery, J. A., Jr.; Frisch, M. J.; Ochterski, J. W.;
(19) Altarawneh, M.; Dlugogorski, B. Z.; Kennedy, E. M.; Mackie, J. Petersson, G. A. A Complete Basis Set Model Chemistry. Vi. Use of
C. Theoretical study of reactions of HO2 in low-temperature Density Functional Geometries and Frequencies. J. Chem. Phys. 1999,
oxidation of benzene. Combust. Flame 2010, 157, 1325−1330. 110, 2822.

11793 DOI: 10.1021/acs.energyfuels.9b02169


Energy Fuels 2019, 33, 11781−11794
Energy & Fuels Article

(36) Siddique, K.; Altarawneh, M.; Gore, J.; Westmoreland, P. R.; (57) Johnson, M. V.; Goldsborough, S. S.; Serinyel, Z.; O’Toole, P.;
Dlugogorski, B. Z. Hydrogen Abstraction from Hydrocarbons by Larkin, E.; O’Malley, G.; Curran, H. J. A Shock Tube Study of n- and
NH2. J. Phys. Chem. A 2017, 121, 2221−2231. iso-Propanol Ignition. Energy.Fuels 2009, 23, 5886−5898.
(37) Simmie, J. M.; Black, G.; Curran, H. J.; Hinde, J. P. Enthalpies (58) Jouzdani, S.; Zhou, A.; Akih-Kumgeh, B. Propanol isomers:
of Formation and Bond Dissociation Energies of Lower Alkyl Investigation of ignition and pyrolysis time scales. Combust. Flame
Hydroperoxides and Related Hydroperoxy and Alkoxy Radicals. J. 2017, 176, 229−244.
Phys. Chem. A 2008, 112, 5010. (59) Zhu, Y.; Davidson, D. F.; Hanson, R. K. 1-Butanol ignition
(38) Somers, K. P.; Simmie, J. M. Benchmarking Compound delay times at low temperatures: An application of the constrained-
Methods (CBS-QB3, CBS-APNO, G3, G4, W1BD) against the Active reaction-volume strategy. Combust. Flame 2014, 161, 634−643.
Thermochemical Tables: Formation Enthalpies of Radicals. J. Phys. (60) Weber, B. W.; Sung, C.-J. Comparative Autoignition Trends in
Chem. A 2015, 119, 8922. Butanol Isomers at Elevated Pressure. Energy.Fuels 2013, 27, 1688−
(39) Canneaux, S.; Bohr, F.; Henon, E. KiSThelP: A Program to 1698.
Predict Thermodynamic Properties and Rate Constants from (61) Tsujimura, T.; Pitz, W. J.; Gillespie, F.; Curran, H. J.; Weber, B.
Quantum Chemistry Results. J. Comput. Chem. 2014, 35, 82. W.; Zhang, Y.; Sung, C.-J. Development of Isopentanol Reaction
(40) Alecu, I. M.; Truhlar, D. G. Computational Study of the Mechanism Reproducing Autoignition Character at High and Low
Reactions of Methanol with the Hydroperoxyl and Methyl Radicals. 2. Temperatures. Energy Fuels 2012, 26, 4871−4886.
Accurate Thermal Rate Constants. J. Phys. Chem. A 2011, 115, (62) da Silva, G.; Bozzelli, J. W.; Liang, L.; Farrell, J. T. Ethanol
14599−14611. Oxidation: Kinetics of the α-Hydroxyethyl Radical + O2 Reaction. J.
(41) Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J. Current Status Phys. Chem. A. 2009, 113, 8923−8933.
of Transition-State Theory. J. Phys. Chem. 1996, 100, 12771. (63) Tang, C.; Wei, L.; Man, X.; Zhang, J.; Huang, Z.; Law, C. K.
(42) Eckart, C. The Penetration of a Potential Barrier by Electrons. High temperature ignition delay times of C5 primary alcohols.
Phys. Rev. 1930, 35, 1303. Combust. Flame 2013, 160, 520−529.
(43) McClurg, R. B.; Flagan, R. C.; Goddard, W. A., III The (64) Togbé, C.; Halter, F.; Foucher, F.; Mounaim-Rousselle, C.;
Hindered Rotor Density-of-States Interpolation Function. J. Chem. Dagaut, P. Experimental and detailed kinetic modeling study of 1-
Phys. 1997, 106, 6675. pentanol oxidation in a JSR and combustion in a bomb. Proc.
(44) Chemkin-Pro 15131. Reaction Design: San Diego, 2013. Combust. Inst. 2011, 33, 367−374.
(45) Koroglu, B.; Pryor, O. M.; Lopez, J.; Nash, L.; Vasu, S. S. Shock (65) Dayma, G.; Togbé, C.; Dagaut, P. Experimental and Detailed
tube ignition delay times and methane time-histories measurements Kinetic Modeling Study of Isoamyl Alcohol (Isopentanol) Oxidation
during excess CO2 diluted oxy-methane combustion. Combust. Flame in a Jet-Stirred Reactor at Elevated Pressure. Energy Fuels 2011, 25,
2016, 164, 152−163. 4986−4998.
(46) Vallabhuni, S. K.; Lele, A. D.; Patel, V.; Lucassen, A.;
Moshammer, K.; AlAbbad, M.; Farooq, A.; Fernandes, R. X.
Autoignition studies of Liquefied Natural Gas (LNG) in a shock
tube and a rapid compression machine. Fuel 2018, 232, 423−430.
(47) Li, J.; Zhao, Z.; Kazakov, A.; Chaos, M.; Dryer, F. L.; Scire, J. J.,
Jr. A comprehensive kinetic mechanism for CO, CH2O, and CH3OH
combustion. Int. J. Chem. Kinet. 2007, 39, 109−136.
(48) Rasmussen, C. L.; Wassard, K. H.; Dam-Johansen, K.; Glarborg,
P. Methanol oxidation in a flow reactor: Implications for the
branching ratio of the CH3OH+OH reaction. Int. J. Chem. Kinet.
2008, 40, 423−441.
(49) Burke, U.; Metcalfe, W. K.; Burke, S. M.; Heufer, K. A.; Dagaut,
P.; Curran, H. J. A detailed chemical kinetic modeling, ignition delay
time and jet-stirred reactor study of methanol oxidation. Combust.
Flame 2016, 165, 125−136.
(50) Noorani, K. E.; Akih-Kumgeh, B.; Bergthorson, J. M.
Comparative High Temperature Shock Tube Ignition of C1−C4
Primary Alcohols. Energy Fuels 2010, 24, 5834−5843.
(51) Millán-Merino, A.; Fernández-Tarrazo, E.; Sánchez-Sanz, M.;
Williams, F. A. A multipurpose reduced mechanism for ethanol
combustion. Combust. Flame 2018, 193, 112−122.
(52) Mittal, G.; Burke, S. M.; Davies, V. A.; Parajuli, B.; Metcalfe, W.
K.; Curran, H. J. Autoignition of ethanol in a rapid compression
machine. Combust. Flame 2014, 161, 1164−1171.
(53) AramcoMech2.0. http://www.nuigalway.ie/
combustionchemistrycentre/mechanismdownloads/aramcomech20/
(30th April, 2019),
(54) Barraza-Botet, C. L.; Wagnon, S. W.; Wooldridge, M. S.
Combustion Chemistry of Ethanol: Ignition and Speciation Studies in
a Rapid Compression Facility. J. Phys. Chem. A. 2016, 120, 7408−
7418.
(55) Metcalfe, W. K.; Burke, S. M.; Ahmed, S. S.; Curran, H. J. A
Hierarchical and Comparative Kinetic Modeling Study of C1− C2
Hydrocarbon and Oxygenated Fuels. Int. J. Chem. Kinet. 2013, 45,
638−675.
(56) Man, X.; Tang, C.; Zhang, J.; Zhang, Y.; Pan, L.; Huang, Z.;
Law, C. K. An experimental and kinetic modeling study of n-propanol
and i-propanol ignition at high temperatures. Combust. Flame 2014,
161, 644−656.

11794 DOI: 10.1021/acs.energyfuels.9b02169


Energy Fuels 2019, 33, 11781−11794

You might also like