You are on page 1of 17

ARTICLE IN PRESS

JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.1 (1-17)

Combustion and Flame ()


www.elsevier.com/locate/combustflame

HCCI experiments with toluene reference fuels modeled by


a semidetailed chemical kinetic model
J.C.G. Andrae a, , T. Brinck b , G.T. Kalghatgi c
a Department of Chemical Engineering and Technology, Royal Institute of Technology (KTH), SE-100 44 Stockholm, Sweden
b Department of Physical Chemistry, Royal Institute of Technology (KTH), SE-100 44 Stockholm, Sweden
c Shell Global Solutions (UK), P.O. Box 1, Chester CH1 3SH, UK

Received 15 February 2008; received in revised form 24 April 2008; accepted 21 May 2008

Abstract
A semidetailed mechanism (137 species and 633 reactions) and new experiments in a homogeneous charge compression ignition (HCCI) engine on the autoignition of toluene reference fuels are presented. Skeletal mechanisms
for isooctane and n-heptane were added to a detailed toluene submechanism. The model shows generally good
agreement with ignition delay times measured in a shock tube and a rapid compression machine and is sensitive
to changes in temperature, pressure, and mixture strength. The addition of reactions involving the formation and
destruction of benzylperoxide radical was crucial to modeling toluene shock tube data. Laminar burning velocities
for benzene and toluene were well predicted by the model after some revision of the high-temperature chemistry.
Moreover, laminar burning velocities of a real gasoline at 353 and 500 K could be predicted by the model using a toluene reference fuel as a surrogate. The model also captures the experimentally observed differences in
combustion phasing of toluene/n-heptane mixtures, compared to a primary reference fuel of the same research
octane number, in HCCI engines as the intake pressure and temperature are changed. For high intake pressures
and low intake temperatures, a sensitivity analysis at the moment of maximum heat release rate shows that the consumption of phenoxy radicals is rate-limiting when a toluene/n-heptane fuel is used, which makes this fuel more
resistant to autoignition than the primary reference fuel. Typical CPU times encountered in zero-dimensional calculations were on the order of seconds and minutes in laminar flame speed calculations. Cross reactions between
benzylperoxy radicals and n-heptane improved the model predictions of shock tube experiments for = 1.0 and
temperatures lower than 800 K for an n-heptane/toluene fuel mixture, but cross reactions had no influence on
HCCI simulations.
2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
Keywords: Toluene reference fuels; Primary reference fuels; Laminar burning velocity; Autoignition; HCCI

1. Introduction

* Corresponding author. Fax: +46 8 696 00 07.

E-mail address: johana@kth.se (J.C.G. Andrae).

Knock, which is a fundamental constraint on the


efficiency of spark ignition (SI) engines, is initiated
by autoignition in the end-gas, the unburned mixture
of fuel and air ahead of the advancing flame front.

0010-2180/$ see front matter 2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2008.05.010
Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
2

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.2 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

In homogeneous charge compression ignition (HCCI)


engines, fuel and air are premixed as in a SI engine,
the mixture is compressed, and heat release occurs by
autoignition of the entire charge. Hence it is of great
interest to understand the autoignition phenomenon in
an internal combustion (IC) engine.
Autoignition depends on the pressure and temperature history of the mixture and on the composition
of the fuel. Fuel autoignition quality, which varies
very widely across different fuels, is a critical property of automotive fuels. It is commonly specified
by research and motor octane numbers, RON and
MON, of the fuel. These are determined by comparing the fuel with primary reference fuels (PRF), mixtures of isooctane and n-heptane, in standard knock
tests. However, practical fuels are very different from
PRF because they are complex mixtures of aromatics,
olefins, paraffins, and oxygenates. They behave quite
differently than PRF in engines as design and operating conditions, which affect the pressure and temperature in the engine, change. The same fuel can match
very different PRFs at different conditions. Thus a
typical European gasoline of 95 RON and 85 MON
will match 95 PRF (95% isooctane) in the RON test
but match 85 PRF in the MON test, but it might be
much more resistant to autoignition than 95 PRF if
tested in a modern SI engine with lower unburnt mixture temperatures at a given pressure than for the RON
test condition. The implications of such observations
are of great practical importance, as discussed in [1].
The difference between the RON and MON of a fuel
is known as sensitivity, S. For PRF, by definition, S is
zero, but for non-PRF fuels, S > 0. The autoignition
behavior of practical fuels, with S > 0, needs to be
understood at a fundamental chemical kinetics level.
Chemical kinetics models have been developed in
the past to explain the autoignition of PRF, but such
schemes are not useful to describe the autoignition
of practical gasolines. On the other hand, practical
fuels are too complex for chemical kinetics models to be developed. However, such models can be
developed for simpler fuel mixtures, surrogates for
gasoline, which are much more like practical gasolines in that they have S > 0. This approach has also
been discussed in [2]. The simplest gasoline surrogate
is a toluene reference fuel (TRF), for which a detailed chemical kinetic scheme has been developed recently [3] using data on ignition delay measurements
from shock tubes [46] and from HCCI engines [7,8].
Detailed chemical kinetics models for other surrogate gasolines [9,10] and ignition delays from shock
tubes [11] have been reported recently. However, the
chemical kinetics model developed for the TRF in [3]
is large (>1000 species) and difficult to use in engine
models. In this paper a semidetailed model is developed that might be easier to incorporate into engine

models. At the same time, we have tried to improve


the chemical mechanism based on new chemical insights.
A semidetailed chemical kinetics model for TRFs
is presented, consisting of 137 species and 633 reactions. Skeleton mechanisms for isooctane and nheptane are added to a detailed mechanism for toluene
oxidation. Model predictions are compared with ignition delay times measured in shock tubes and in
a rapid compression machine for neat fuels as well
as for fuel mixtures over practical ranges of temperature and pressure. The model has also been validated
against experiments for laminar burning velocities of
benzene/air, toluene/air, and a real gasoline/air mixture. New experiments with toluene reference fuels
have been conducted in a HCCI engine under different operating conditions and model predictions are
compared with those and other HCCI experiments to
test the capability to correlate ignition delays to empirical measures of autoignition quality of a fuel under different operating conditions.

2. Construction of kinetic model


The methodology was to have a detailed description of toluene chemistry as base model with complementary skeletal mechanisms for primary reference
fuels isooctane and n-heptane. The starting mechanism for the toluene chemistry was a kinetic model for
benzene from Alzueta et al. [12], developed against
flow-reactor conditions at excess air ratios ranging
from close to stoichiometric to very lean, temperature range 9001450 K, and residence times on the
order of 150 ms. To this validated mechanism for
benzene were added reactions for toluene oxidation
based on the work of Sivaramakrishnan et al. [13]
and our recent work with a detailed mechanism for
TRF [3]. Some additional reactions involving the
phenyl methyl radical (C6 H4 CH3 ) were added, based
on the work by Bounaceur et al. [14]. The toluene submechanism is shown in Appendix A.
In [3] it was found necessary to include a global
branching reaction involving benzyl radicals and oxygen in order to increase reactivity in the toluene mechanism at high pressure. Here we have tried to improve
that description with elementary reactions including
formation of the benzylperoxy radical [15,16] (the
numbers in parentheses indicate the reaction number
as listed in Appendix A),
C6 H5 CH2 + O2 C6 H5 CH2 OO,

(33)

followed by decomposition of the benzylperoxy radical to benzaldehyde and the hydroxyl radical,
C6 H5 CH2 OO C6 H5 CHO + OH.

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

(34)

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.3 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
3

The rate constants for (33) were taken from the work
of Fenter et al. [17]. For reaction (34) the rate constant was tuned for best fit to shock tube experiments
with toluene/air to have a value between the ones proposed by Bounaceur et al. [14] and Lindstedt and
Maurice [18].
The toluene mechanism has been complemented
with skeleton mechanisms for n-heptane [19] and
isooctane [20] and these submechanisms are shown
in Appendixes B and C, respectively.
The skeletal n-heptane mechanism describes both
the high- and the low-temperature chemistry and has
been shown to reproduce ignition delay times at various pressures and temperatures reasonably well. The
skeletal isooctane mechanism by Tanaka et al. [20],
developed especially for HCCI conditions, has indicated good performance over wide ranges of pressure,
temperature, and equivalence ratios, especially at high
pressures and lean equivalence ratios.
Cross reactions were added to the TRF mechanism based on our previous work [3,21]. In this work
we propose that cross reactions between benzylperoxy radicals and the primary reference fuels may also
be of importance. These reactions involve hydrogen
abstraction from primary and secondary carbons of
n-heptane. Since reliable experimental data for the kinetics do not exist, we have analyzed the reactions
using modern quantum chemical methods based on
KohnSham density functional theory (DFT). Optimized geometries of reactants, products, and transition states have been obtained at the B3LYP/6-31+G*
level with the Gaussian 03 suite of programs [22].
Accurate energies for the optimized structures have
been determined at the B2PLYP/cc-pVTZ level using
the Orca program package [23]. The B2PLYP//ccpVTZ level is expected to produce activation barriers for the studied reactions that are accurate to
within ca. 3 kcal/mol [2325]. The temperature dependence of the derived rate constants has been determined from the B3LYP/6-31+G* vibrational frequencies. Anharmonicity in the vibrations due to hindered
internal rotation was estimated using a locally developed approach [26]. On the basis of the computations,
temperature-dependent rate constants were estimated
for the temperature interval of 3001100 K.
The overall mechanism consists of 137 species
and 633 reactions. Table 1 shows an overview of
species included. Only around 15% of the species
from the mechanism describing the primary reference
fuels are used, making the mechanism very compact.
The full mechanism in CHEMKIN format with associated thermodynamic and transport data is provided
as supplementary material to the article.
The next section describes the comparison between experimental values of ignition delay times in
shock tubes and rapid compression machine as well

as for laminar burning velocities and numerical simulations using the chemical kinetic model.

3. Model validation
3.1. Comparison with ignition delay times in shock
tubes
A parameter often used in autoignition studies is
the autoignition delay, , which is the time required
for autoignition to happen once the fuel/air mixture is
raised to a given pressure and temperature and held
at that condition in a rapid compression machine or
a shock tube. In this work, model predictions have
been compared with shock tube autoignition delay
time data over practical ranges of temperature and
pressure.
The Chemkin Collection [27] has been employed
to handle all kinetic and thermodynamic data, as well
as for simulations. The closed homogeneous constant
volume reactor model (AURORA application) has
been used to simulate the conditions behind the reflected shocks (p5 , T5 ). In shock tube experiments,
if there is a small change in pressure (generally resulting from an energy release in the reacting mixture), the shock tube flow can follow this pressure
rise. The pressure rise is dependent on how much
energy is released. If very little energy is released,
then the constant volume and constant pressure constraints give approximately the same ignition times.
But for mixtures with higher fuel concentrations, the
constant volume constraint gives shorter and more accurate ignition times. What actually happens in the
shock tube is that as the reacting gas heats up and
the pressure rises, about half the maximum pressure
rise can be achieved/maintained and the other half
of the energy goes into expanding the gas volume.
If the pressure rise is very rapid, then the constant
volume method is a very good approximation, up to
the point of ignition. If the pressure rise is slower,
the constant volume assumption is still more accurate [28]. Autoignition delay times, , were defined
as the times needed for the mixture to reach a temperature of 400 K above the initial temperature. This
definition was close to the time for the temperature inflection point, = t (dT /dt)max , as determined with
the AURORA program [27].
3.1.1. Primary reference fuels
Fig. 1 shows the results when modeled results for
primary reference fuels are compared with the ones
measured by Fieweger et al. [29]. The model is sensitive to changes in both temperature and mixture composition and captures the negative temperature coefficient (NTC) region. Some tuning and some addition

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
4

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.4 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Table 1
List of species included in the mechanism
C6 H5 CH3
C6 H5 CH2
OC6 H4 CH3
HOC6 H4 CH3
C6 H4 CH3
C6 H5 CH2 OO
C6 H5 CH2 OOH
C6 H5 CH2 O
C6 H5 CHO
Bibenzyl
C6 H5 CO
C6 H5 CHCH2
C6 H5 C2 H4
C6 H5 C2 H5
C6 H5 CH2 OH
CO
CO2
C6 H6
C6 H5
C6 H4
C6 H5 O
C6 H5 OH
OC6 H4 O
oOC6 H4 O
OC6 H3 O
C6 H3 O3
C5 H7
C5 H6
C5 H5
C5 H5 O
C5 H5 OH
C5 H4 O
C5 H4 OH
C5 H4
C5 H3 O
C5 H3
CH2 CHCHCHCH3
CH2 CHC.HCHCH2
CH2 CHCHCHC.H2
CHCCHCHC.H2
CH2 CHC.HCHCHOH
CH2 CHCHCHCH2 OH
CHOCH2 CH2 C.HCHO
CH2 CHCHCH2

CH2 CHC.CH2
CH2 CHCHC.H
CH2 CHCCH
HCCHCCH
H2 CCCCH
C4 H2
C4 H
CHOCH2 CH2 CHO
CHOCH2 CH2 C.O
CH2 CHC.HCHO
CH2 CHCHCO
C.HCHCHCO
CH2 C.CHCO
CHCCHCO
CHCC.CO
H2 C4 O
CH2 CHCH3
CH2 CHC.H2
CH2 C.CH3
C.HCHCH3
H2 CCCH2
H3 CCCH
H2 CCCH
C3 H2
CH2 CHCHO
CHOCHCHOH
C.HCHCHO
CH2 CHC.O
C2 H6
C2 H5
C2 H4
C2 H3
C2 H2
C2 H
C2
CH3 HCO
CH3 CO
CH2 HCO
C2 H2 OH
OCHCHO
CH2 CO
HCCOH
HCCO
C2 O

of reactions were needed on the skeletal mechanism


for isooctane by Tanaka et al. [20] to improve the
results (see Appendix C). Tanaka et al. concentrated
on lean fuel/air mixtures and did not compare their
proposed model with results involving stoichiometric
mixtures of isooctane and air.
However, no changes were needed in the skeletal
n-heptane mechanism upon merging with the detailed
toluene model. When modeled ignition delay times
for stoichiometric n-heptane/air mixtures were compared with the experimental results by Gauthier et
al. [6], the average relative error was only around 6%.

O2 CCHOO.
C2 H5 CHO
C2 H5 CO
CH4
CH3
CH2
CH2 (S)
CH
C
CH3 OH
CH3 O
CH2 OH
CH2 O
HCO
HCOO
H
O
OH
H2
O2
HO2
H2 O
H 2 O2
NO
HNO
HONO
NO2
N2
C7 H16
C7 H15 -1
C7 H15 -2
C7 H15 O2
C7 H14 O2 H
O2 C7 H14 O2 H
HO2 C7 H13 O2 H
OC7 H13 O2 H
OC7 H13 O
C5 H11
I C3 H7
PC4 H9
C8 H18
AC8 H17
AC8 H17 OO
AC8 H16 OOHB

AC8 H16 OOHBOO


OC8 H15 OOH
OC8 H15 O
JC8 H16
IC4 H8

This is in strong contrast to the detailed model presented in [3], where the modeled values were a factor
of 2 higher than experiments for some temperatures.
This indicates that the skeletal n-heptane mechanism
used in this work (not a reduced mechanism of the
detailed n-heptane model in [3]) is better tuned to
predict ignition delay times in shock tubes. Moreover, Herzler et al. [30] measured autoignition delay
times in a shock tube for lean n-heptane/air mixtures
( = 0.10.4) at p = 5.0 MPa and Fig. 2 shows modeled values compared with experiments for = 0.2
and 0.3. There is good agreement with the proposed

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.5 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
5

Fig. 1. Comparison between experimental shock tube ignition delay data for primary reference fuels from Fieweger
et al. [29] and numerical simulations using the semidetailed
chemical kinetic model. Conditions behind reflected shock
(p5 , T5 ) simulated as a closed homogeneous constant volume adiabatic reactor.

Fig. 3. Comparison between experimental shock tube ignition delay data from Davidson et al. [5] and numerical
simulations using the semidetailed chemical kinetic model.
Conditions behind reflected shock (p5 , T5 ) simulated as a
closed homogeneous constant volume adiabatic reactor. Experimental data normalized with exp = f (T )p0.93 .

Fig. 2. Comparison between experimental shock tube ignition delay data from Herzler et al. [30] and numerical
simulations using the semidetailed chemical kinetic model.
Conditions behind reflected shock (p5 , T5 ) simulated as a
closed homogeneous constant volume adiabatic reactor.

Fig. 4. Comparison between experimental shock tube ignition delay data from Davidson et al. [5] and numerical
simulations using the semidetailed chemical kinetic model.
Conditions behind reflected shock (p5 , T5 ) simulated as a
closed homogeneous constant volume adiabatic reactor. Experimental data normalized with exp = f (T )p0.5 .

model except at temperatures below around 750 K.


This is in strong contrast to the detailed model for nheptane used in [3], where the modeled values have
been found to be systematically a factor of 2 higher
than experiments [30].
3.1.2. Toluene
Figs. 3 and 4 show the results when modeled results for toluene are compared with the ones measured
by Davidson et al. [5], and generally the model captures the change in for changes in temperature and
mixture strength. Reaction (34) was found to be very
sensitive to the overall reactivity for neat toluene/air
mixtures and the rate constant was tuned for best fit
of model to the shock tube data to have a value be-

tween the one proposed by Bounaceur et al. [14] and


Lindstedt and Maurice [18].
3.1.3. Toluene reference fuels
The next step is to compare the model predictions
with data for toluene reference fuels. It is not obvious
that a model that shows good agreement with data for
single-component fuels gives good agreement for fuel
mixtures. In [3] we showed that
C6 H5 CH2 + HO2 C6 H5 CH2 O + OH,

(27)

whereby two unreactive radicals are transformed to


more reactive ones, was very sensitive to crossacceleration effects for toluene/n-heptane fuel mixtures, not seen for the single fuels, and the rate con-

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
6

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.6 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Table 2
Composition (liquid volume %) of toluene reference fuels investigated and their autoignition properties
Fuel

n-Heptane

Isooctane

Toluene

RON

MON

Ref.

Surr. A
Surr. B
TOLHEP
TOLHEP1
TOLHEP3

17
17
35
36
50

63
69

20
14
65
64
50

88
87
83.9
82.3
64.1

85
85
73.2
73.1
58.1

[6]
[6]
[7,21]
[8]
[8]

Fig. 5. Comparison between experimental shock tube ignition delay data from Herzler et al. [4] and numerical
simulations using the semidetailed chemical kinetic model.
Conditions behind reflected shock (p5 , T5 ) simulated as a
closed homogeneous constant volume adiabatic reactor. Experimental data normalized with exp = f (T )p 0.883 .

stant had to be reduced compared to previous models


to improve model predictions. For the same reason,
it has been shown that reaction (27) would be the
cause for antagonistic blending octane numbers for
mixtures with isooctane and toluene [9].
Figs. 5 and 6 show the results when modeled results for a toluene/n-heptane fuel (TOLHEP in Table 2) are compared with the ones measured by Herzler et al. [4]. The semidetailed kinetic model captures the change in for changes in temperature and
pressure satisfactorily. The reaction with phenol and
oxygen to give phenoxy and hydroperoxy radicals,
C6 H5 OH + O2 C6 H5 O + HO2 ,

(88)

was found to be very sensitive under lean conditions ( = 0.3), with the estimated rate given by
Alzueta et al. [12] overpredicting compared to experiments. In order to have better predictions, the
rate for the forward reaction was lowered from k88 =
1.0 1013 exp(38 kcal mol1 /RT ) cm3 mol1 s1
to k88 = 8.0 1012 exp(40 kcal mol1 /RT ) cm3
mol1 s1 . For = 1.0 the model is somewhat too
reactive at 5.0 MPa, and also, at temperatures lower
than around 820 K, the model predictions start to deviate from experiments (see Fig. 6). The introduction

Fig. 6. Comparison between experimental shock tube ignition delay data from Herzler et al. [4] and numerical
simulations using the semidetailed chemical kinetic model.
Conditions behind reflected shock (p5 , T5 ) simulated as a
closed homogeneous constant volume adiabatic reactor. Experimental data normalized with exp = f (T )p1.06 .

of cross reactions between benzylperoxy radicals and


n-heptane,
C6 H5 CH2 OO + C7 H16
C6 H5 CH2 OOH + C7 H15 -1,

(619)

C6 H5 CH2 OO + C7 H16
C6 H5 CH2 OOH + C7 H15 -2,

(620)

improves the predictions at temperatures below


around 800 K, as illustrated for p = 3.0 MPa in Fig. 6.
The estimated activation barrier based on quantum
chemical methods was 17.6 3 kcal/mol for abstraction of secondary hydrogen atoms from n-heptane by
benzylperoxy radicals.
Figs. 7 and 8 show the results when modeled results for two gasoline surrogate fuels (Surrogates A
and B in Table 2) are compared with the ones measured by Gauthier et al. [6]. Generally good agreement is obtained, although it has to be said that no
shock tube experiments are available at lower temperatures for these two fuels to compare with model
predictions. However, we have performed new experiments in the Shell HCCI engine at Thornton, which
are presented in Section 4.2 of this paper.
Modeled results having been compared with experimental shock tube ignition delays at practical

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.7 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
7

Fig. 7. Comparison between experimental shock tube ignition delay data for Surrogate A from Gauthier et al. [6]
and numerical simulations using the semidetailed chemical kinetic model. Conditions behind reflected shock
(p5 , T5 ) simulated as a closed homogeneous constant volume adiabatic reactor. Experimental data normalized with
exp = f (T )p0.83 .

Fig. 9. Comparison of experimental [31] and modeling results (lines) for PRFs and TSFs in a rapid compression machine (RCM). T = 318 K, p = 0.1 MPa, compression ratio
16. A heat loss of 15 cal/s was assumed while keeping the
volume constant after reaching top dead center in the RCM
modeling.

shown in Fig. 9. Generally the model does a very good


job at predicting the experiments and further validates
the semidetailed model to simulate toluene reference
fuels at HCCI conditions.
3.3. Laminar burning velocities

Fig. 8. Comparison between experimental shock tube ignition delay data for Surrogate B from Gauthier et al. [6]
and numerical simulations using the semidetailed chemical kinetic model. Conditions behind reflected shock
(p5 , T5 ) simulated as a closed homogeneous constant volume adiabatic reactor. Experimental data normalized with
exp = f (T )p0.96 .

ranges of pressure and temperature for single-component fuels as well as for fuel mixtures with TRFs,
the model can be said to show very good agreement
at high temperatures and in the NTC region, but is
weaker at temperatures below 750 K.
3.2. Comparison with ignition delay times in a rapid
compression machine
In order to validate the model against rapid compression machine (RCM) data, we have simulated
the experiments carried out by Tanaka et al. [31] for
primary and toluene reference fuels. The results are

The laminar burning velocity is a characteristic response of a given combustible mixture and embodies
the fundamental diffusive, reactive, and exothermic
mixture properties. It is a common target used to partially validate kinetic mechanisms. Moreover, flame
speed is generally important in engine combustion.
Therefore the semidetailed model was also
checked for validity by comparing model predictions
against experimental values of laminar burning velocities for benzene/air and toluene/air mixtures by Davis
et al. [32] and a real gasoline/air mixture by Zhao et
al. [33]. The calculations were performed with the
PREMIX program of the Chemkin Collection [27]
and a number of dipole moments and polarizabilities in the transport data base were updated with data
from the Handbook of Chemistry and Physics [34].
A first comparison showed that the model predicted
around 15 cm/s lower burning velocities for benzene and toluene than the experiments. That indicated that revisions had to be made, and based on
a flow rate A-factor sensitivity analysis, some rate
constants in the high-temperature chemistry of the
benzene submechanism were changed in order to improve the predictions. It was also necessary to add
a reaction for the high-temperature decomposition
of isobutene (iC4 H8 ), not important in the autoignition delay time calculations in shock tubes and RCM
above. Some additional reactions were also added for

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
8

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.8 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

the peak burning velocity than the PRF 87. Overall,


the laminar flame speed simulations have shown that
the model is sensitive to molecular structure effects
on laminar flame speeds.

4. Homogeneous charge compression ignition


engine experiments and modeling

Fig. 10. Comparison of experimental [32] and computed


laminar burning velocities for benzene/air and toluene/air
mixtures at room temperature and 1 atm. The accuracy of
the experiments was within 12 cm/s.

Fig. 11. Comparison of experimental laminar burning velocities (filled symbols) for a full boiling gasoline fuel/air mixture [33] and computed values for surrogate gasoline fuels in
air. Solid lines and open symbolsSurrogate B in Table 2.
Dashed lines and open symbolsPRF 87. p = 0.1 MPa.

ethylbenzene. Appendix E summarizes the changes


made in the mechanism in order to simulate laminar
burning velocities. Fig. 10 shows the results for benzene/air and toluene/air mixtures. After the revision,
the model generally does a good job of predicting
the laminar burning velocities of both benzene and
toluene although the experiments are underpredicted
by the model for < 1.1. Zhao et al. [33] measured
laminar burning velocities at high temperatures for a
real gasoline fuel, CR-87. CR-87 has octane number
87 and contains more than one hundred components,
and the relative concentration of any single component is less than 2%. In the simulations, two different
surrogates for gasoline to represent the real gasoline
were tested, PRF 87 and Surrogate B in Table 2, that
has a RON of 87. Fig. 11 shows the result of the calculations and the model is able to predict the laminar
burning velocity as a function of temperature and .
The toluene reference fuel gives better prediction for

In this section the model will be tested against


homogeneous charge compression ignition engine experiments, a typical target application, as the autoignition process in HCCI engines is driven by chemical
kinetics.
In practice, compared to a primary reference fuel,
a sensitive fuel (with aromatics, olefins, oxygenates)
becomes more resistant to autoignition if the pressure
is increased for a given temperature and less resistant to autoignition if the temperature is increased for
a given pressure. For all practical purposes, for both
knock and HCCI, the true autoignition quality of a
non-PRF fuel is given by the octane index [1] OI =
RON KS, where sensitivity S = RON MON. K is
constant depending on operating conditions and is
positive or negative depending on whether the temperature for a given pressure is higher or lower than
the condition where the autoignition quality is determined by RON (K = 0). A single-zone engine
model (no crevices and charge inhomogeneities) has
been used to evaluate how well the validated mechanism could capture autoignition behavior for conditions corresponding to HCCI engine combustion. The
usefulness of the single-zone assumption is in providing an estimate of ignition delay time as a function of
thermodynamic conditions in the combustion chamber.
4.1. HCCI engine experiments at KTH
Fig. 12 shows model predictions (single zone
model) and average pressure and heat release rate
based on 100 engine cycles from the HCCI engine
at KTH [7] for the TOLHEP fuel in Table 2 validated
against shock tube ignition delay data as shown in
Figs. 56, and a PRF with RON = MON = 84. The
conditions for the calculations are as OP1 in Table 3
with the heat transfer model the same as in [3]. According to the experiment, the two fuels show the
same resistance to autoignition under this particular
operating condition, where K is +0.07 [7] and the
model is able to predict this as well. Fig. 13 shows
model predictions for OP2 in Table 3 and average
experimental results when the K value is 1.5 at a
higher intake pressure and lower intake temperature.
The octane index for the toluene reference fuel is by

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.9 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
9

Fig. 12. Experimental and calculated pressure and heat release rate as functions of crank angle in the HCCI engine at
KTH. OP1 in Table 3. Modeled heat release rates divided by
a factor of 10. Initial temperature and pressure at the start
of calculations (99 crank angle before top dead center) are
447 K and 1.7 bar, respectively.

Fig. 14. Experimental and calculated pressures as a function of crank angle in the HCCI engine at Shell. OP3 in
Table 3. Initial temperature and pressure at the start of calculations (99 crank angle before top dead center) are 489 K
and 1.6 bar, respectively.

fecting the pressure and temperature development in


the engine change.
4.2. HCCI engine experiments at Shell Technology
Centre Thornton

Fig. 13. Experimental and calculated pressure and heat release rate as functions of crank angle in the HCCI engine
at KTH. OP2 in Table 3. Modeled heat release rate divided
by a factor of 5. Initial temperature and pressure at the start
of calculations (99 crank angle before top dead center) are
421 K and 3.2 bar, respectively.

definition 101 and would therefore have higher resistance to autoignition than PRF 84. Again, the model is
able to predict this trend. Thus the model can capture
the different autoignition behavior of paraffinic and
nonparaffinic fuels as the experimental conditions af-

Fig. 14 shows model predictions (single zone


model) and average pressure based on 100 engine
cycles from the HCCI engine at Shell Technology
Centre Thornton [8] for TOLHEP1 and TOLHEP3 in
Table 2 and OP3 in Table 3. The model does a good
job of differentiating the two fuels according to their
resistance to autoignition at high intake temperature
and low intake pressure (K was +2.2 in this case).
For this study, new HCCI experiments have also
been conducted for Surrogate A and B (see Table 2) in
the HCCI engine at Thornton, i.e., the same fuels validated against shock tube ignition delay data as shown
in Figs. 7 and 8 above. The HCCI engine is described
in more detail in [8] and the experimental data from
the engine tests (measured average cylinder pressure
as a function of crank angle from 100 engine cycles)
for two different operating conditions (OP4 and OP5
in Table 3) is given as supplementary data to the paper. Figs. 15 and 16 show the results for Surrogate B,
first where the intake temperature is high and intake

Table 3
HCCI operating conditions

OP1
OP2
OP3
OP4
OP5

pin (MPa)

Tin ( C)

Speed (rpm)

Ref.

0.2857
0.25
0.2857
0.25
0.25

0.1
0.2
0.1
0.1
0.2

120
40
250
250
80

900
900
1200
1200
1200

[7,21]
[7,21]
[8]
This work
This work

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
10

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.10 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Fig. 15. Experimental and calculated pressures for Surrogate


B as a function of crank angle in the HCCI engine at Shell.
OP4 in Table 3. Initial temperature and pressure at the start
of calculations (99 crank angle before top dead center) are
489 K and 1.5 bar, respectively.

Fig. 17. Normalized sensitivity coefficients for the 10 most


sensitive reactions at the crank angle for maximum heat release rate (2.3 CA ATDC). TOLHEP fuel and initial conditions as in Fig. 13.

5. Discussion

Fig. 16. Experimental and calculated pressures for Surrogate


B as a function of crank angle in the HCCI engine at Shell.
OP5 in Table 3. Initial temperature and pressure at the start
of calculations (99 crank angle before top dead center) are
435 K and 3.3 bar, respectively.

pressure low (Fig. 15) and second where the intake


temperature is low and intake pressure high (Fig. 16),
i.e. operating conditions are changing from a highly
positive K value of +2.2 to a strongly negative value
of 1.5. Together with the experimental results are
shown model predictions with the semidetailed model
in this work and the detailed model in [3]. Typical
CPU times for the semidetailed model were on the
order of seconds and for the detailed model (>1000
species) on, the order of minutes. For both cases the
detailed model is much less reactive than the semidetailed model for the same initial conditions. This may
in part be explained by the results presented in Section 3.1.1.

It was seen in Figs. 12 and 13 that the model


was able to predict the shift in resistance to autoignition for an n-heptane/toluene fuel compared to a PRF
when the operating conditions changed in a HCCI engine from having high intake temperature and low
intake pressure to low intake temperature and 1 bar
boost intake pressure. In [21] the introduction of cross
reactions into a detailed model for TRF was found
to improve the results in those particular HCCI simulations. However, the detailed model in [21] had
not been compared with shock tube data under conditions relevant to HCCI combustion and the rates
assigned to those cross reactions were too high [3].
To check the importance of cross reactions in the
current semidetailed model, we performed simulations with and without the reactions in Appendix D,
and no difference could be seen. Therefore we conclude that there must be some other explanation in
the chemistry that would be responsible for the difference in autoignition when changing the operating
conditions. Figs. 17 and 18 show normalized sensitivity coefficients for the n-heptane/toluene fuel as
calculated with Chemkin for the 10 most sensitive
reactions at the crank angle for maximum heat release rate. The conditions for the calculations are as in
Figs. 12 and 13. For the case with high intake pressure
(see Fig. 17), the reaction with the highest positive
sensitivity (increased rate leads to decreased ignition
delay) is the ring-opening reaction involving the phenoxy radical,
C6 H5 O C5 H5 + CO,

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

(78)

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.11 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
11

Fig. 18. Normalized sensitivity coefficients for the 10 most


sensitive reactions at the crank angle for maximum heat release rate (4.3 CA ATDC). TOLHEP fuel and initial conditions as in Fig. 12.

Fig. 19. Predicted temporal mole fraction profiles for phenol (C6 H5 OH) and 2,4-cyclopentadiene-1-one (C5 H4 O) for
OP1 and OP2 in Table 3 (see Figs. 10 and 13). Fuel is TOLHEP in Table 2.

with rate constant k78 = 7.4 1011 exp(43.9 kcal


mol1 /RT ) s1 . The reaction to form phenoxy from
phenol in
C6 H5 OH + C2 H3 C6 H5 O + C2 H4

(90)

with rate constant k90 = 6.0 1012 cm3 mol1 s1 ,


and the one with 2,4-cyclopentadiene-1-one (C5 H4 O)
in
C5 H4 O + OH C5 H3 O + H2 O,

(171)

with rate constant k171 = 1.1 108 T 1.4 exp(1.45


kcal mol1 /RT ) cm3 mol1 s1 also shows a high
positive sensitivity.
For the case of high intake temperature and low
intake pressure, on the other hand, reactions (78),
(90), and (171) are not among the 10 most sensitive reactions. Instead it is the decomposition of benzylperoxide (34) that has the highest positive sensitivity. Fig. 19 shows temporal profiles of mole fractions for phenol (C6 H5 OH) and 2,4-cyclopentadiene1-one (C5 H4 O) for the two operating conditions. The
rates of disappearance of C6 H5 OH and C5 H4 O are
slower at the moment of autoignition for the case with
boosted pressure (2 bar) compared to the case with
low intake pressure (1 bar), implying that reactions
(78), (90), and (171) are rate-limiting when increasing
the pressure and decreasing the intake temperature.
For both operating conditions the reaction
C6 H5 OH + O2 C6 H5 O + HO2

(88)

showed the highest negative sensitivity (increased


rate giving longer ignition delay). Reaction (88) also
showed a strong negative sensitivity for lean conditions ( = 0.3) in shock tube simulations for the same
n-heptane/toluene fuel mixture as used in the HCCI
experiments and modeling.

Fig. 20. Normalized sensitivity coefficients for the 10 most


sensitive reactions at the crank angle for maximum heat release rate (9.0 CA ATDC). PRF 84 fuel and initial conditions as in Fig. 13.

In the case with low intake pressure and high intake temperature, the consumption of phenoxy radicals is not rate-limiting in the same way as for the
case with high intake pressure and low intake temperature.
Figs. 20 and 21 show normalized sensitivity coefficients for fuel PRF 84 as calculated with Chemkin
for the 10 most sensitive reactions at the crank angle for maximum heat release rate. For both operating
conditions the reaction
AC8 H16 OOHB + O2
AC8 H16 OOHBOO

(605)

shows the highest positive sensitivity and the reaction


AC8 H17 + O2 JC8 H16 + HO2

(609)

shows the highest negative sensitivity. Although there


are some small variations, the trend in sensitivity to

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
12

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.12 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Fig. 21. Normalized sensitivity coefficients for the 10 most


sensitive reactions at the crank angle for maximum heat release rate (4.3 CA ATDC). PRF 84 fuel and initial conditions as in Fig. 12.

reactions is more similar for the PRF with changing


operating conditions than for the n-heptane/toluene
fuel (i.e., the non-PRF). This would indicate that
it is the changes in the aromatic chemistry in the
n-heptane/toluene fuel that are responsible for the
changes in resistance to autoignition for different operating conditions.
The semidetailed model for TRF developed in this
work against ignition delay data in shock tube and
rapid compression machine at practical conditions for
pressure and temperature as well as laminar burning
velocities has been shown to be predictive in terms of
autoignition phasing when tested in single zone calculations (0-D) against experimental data for PRF and
TRF in two different HCCI engines and under different operating conditions. This makes us confident
that the model would be very useful in more complex
engine simulations. The CPU times encountered are
in the order of seconds in zero-dimensional calculations and minutes in laminar flame speed simulations
(1-D).

6. Conclusions
A semidetailed mechanism (137 species and 633
reactions) and new experiments in a HCCI engine on
the autoignition of toluene reference fuels have been
presented.
The model shows generally good agreement when
compared to ignition delay times measured in shock

tubes and rapid compression machines and is sensitive to changes in temperature, pressure, and mixture
strength.
The addition of reactions involving the formation
and destruction of benzylperoxide radicals was crucial to modeling toluene shock tube data.
Cross reactions between benzylperoxy radicals
and n-heptane improved the model predictions at temperatures lower than 800 K for an n-heptane/toluene
fuel mixture.
Laminar burning velocities for benzene, toluene,
and a real gasoline surrogate fuel were well predicted by the model after some revision of the hightemperature chemistry, and the model is sensitive to
molecular structure effects on laminar flame speeds.
In a HCCI engine, the model can predict the shift
in resistance in autoignition for an n-heptane/toluene
fuel compared to a PRF when the operating conditions change from low intake pressure/high intake
temperature to high intake pressure/low intake temperature. A sensitivity analysis at the moment of maximum heat release rate shows that the consumption
of phenoxy radicals is rate-limiting for high intake
pressure case when a toluene/n-heptane fuel is used,
which makes this fuel more resistant to autoignition
than the primary reference fuel.
The new model can also differentiate different nheptane/toluene fuels at high intake temperature and
low intake pressure and shows higher reactivity for
a fuel mixture of isooctane/toluene/n-heptane when
compared with a detailed model for TRF (>1000
species). This would be partly explained by a slower
reactivity of the n-heptane submechanism in the detailed model at lean fuel/air ratios.
Acknowledgments
Shell Global Solutions (UK) is acknowledged for
financially supporting this work. Bob Head played an
important part in performing the HCCI experiments
at Shell Technology Centre Thornton.
Supplementary material
The online version of this article contains additional supplementary material.
Please visit DOI: 10.1016/j.combustflame.2008.
05.010.

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.13 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
13

Appendix A
Table A
Toluene subset with rate constants

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53

C6 H5 CH2 + H C6 H5 CH3
C6 H5 CH3 C6 H5 + CH3
C6 H5 CH3 + O2 C6 H5 CH2 + HO2
C6 H5 CH3 + OH C6 H5 CH2 + H2 O
C6 H5 CH3 + O C6 H5 CH2 + OH
C6 H5 CH3 + HO2 C6 H5 CH2 + H2 O2
C6 H5 CH3 + H C6 H5 CH2 + H2
C6 H5 CH3 + H C6 H6 + CH3
C6 H5 CH3 + O OC6 H4 CH3 + H
CH3 + C6 H5 CH3 CH4 + C6 H5 CH2
C6 H5 + C6 H5 CH3 C6 H6 + C6 H5 CH2
C6 H5 CH3 + H C6 H4 CH3 + H2
C6 H5 CH3 + O C6 H4 CH3 + OH
C6 H5 CH3 + OH C6 H4 CH3 + H2 O
C6 H5 CH3 + HO2 C6 H4 CH3 + H2 O2
C6 H5 CH3 + CH3 C6 H4 CH3 + CH4
C6 H4 CH3 + O2 OC6 H4 CH3 + O
C6 H4 CH3 + H C6 H5 CH3
C6 H4 CH3 + H C6 H5 CH2 + H
C6 H4 CH3 + O OC6 H4 CH3
C6 H4 CH3 + OH HOC6 H4 CH3
C6 H4 CH3 + HO2 OC6 H4 CH3 + OH
C6 H5 CH2 C5 H5 + C2 H2
C6 H5 CH2 H2 CCCH + CH2 CHCCH
C6 H5 CH2 + O C6 H5 CHO + H
C6 H5 CH2 + O C6 H5 + CH2 O
C6 H5 CH2 + HO2 C6 H5 CH2 O + OH
C6 H5 CH2 O C6 H5 CHO + H
C6 H5 CH2 O C6 H5 + CH2 O
C6 H5 CH2 + C6 H5 CH2 Bibenzyl
C6 H5 C2 H5 C6 H5 CH2 + CH3
C6 H5 CH2 + OH C6 H5 CH2 OH
C6 H5 CH2 + O2 C6 H5 CH2 OO
Reverse Arrhenius coefficients:
C6 H5 CH2 OO C6 H5 CHO + OH
C6 H5 CH2 + O2 C6 H5 CH2 O + O
C6 H5 CHO + O2 C6 H5 CO + HO2
C6 H5 CHO + OH C6 H5 CO + H2 O
C6 H5 CHO + H C6 H5 CO + H2
C6 H5 CHO + H C6 H6 + HCO
C6 H5 CHO + O C6 H5 CO + OH
C6 H5 CH2 + C6 H5 CHO C6 H5 CH3 + C6 H5 CO
CH3 + C6 H5 CHO CH4 + C6 H5 CO
C6 H5 + C6 H5 CHO C6 H6 + C6 H5 CO
C6 H5 CO C6 H5 + CO
C6 H5 CH2 OH + O2 C6 H5 CHO + HO2 + H
C6 H5 CH2 OH + OH C6 H5 CHO + H2 O + H
C6 H5 CH2 OH + H C6 H5 CHO + H2 + H
C6 H5 CH2 OH + H C6 H6 + CH2 OH
C6 H5 CH2 OH + C6 H5 CH2 C6 H5 CHO + C6 H5 CH3 + H
C6 H5 CH2 OH + C6 H5 C6 H5 CHO + C6 H6 + H
C6 H5 C2 H5 + OH C6 H5 C2 H4 + H2 O
C6 H5 C2 H5 + HO2 C6 H5 C2 H4 + H2 O2
C6 H5 C2 H5 + O C6 H5 C2 H4 + OH

Aa

na

1.80E+14
1.40E+16
1.81E+12
6.70E+09
6.30E+11
1.30E+11
6.00E+13
1.20E+13
1.63E+13
3.16E+11
7.94E+13
6.00E+08
2.00E+13
3.40E+08
4.00E+11
2.00E+12
2.60E+13
1.00E+14
1.00E+13
1.00E+14
1.00E+13
5.00E+12
6.00E+13
2.00E+14
2.50E+14
8.00E+13
2.00E+12
2.00E+13
2.00E+13
2.51E+11
2.00E+15
6.00E+13
4.60E+11
4.38E+13
3.00E+09
6.32E+12
1.02E+13
1.71E+09
5.00E+13
1.20E+13
9.04E+12
2.77E+03
2.77E+03
7.01E+11
3.98E+14
2.00E+14
8.43E+12
8.00E+13
1.20E+13
2.11E+11
1.40E+12
4.80E+12
6.20E+04
2.23E+13

0
0
0
1
0
0
0
0
0
0
0
1
0
1.4
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.4
0
0
0
0
1.1
0
0
1.2
0
0
0
2.8
2.8
0
0
0
0
0
0
0
0
0
2.5
0

Ea

Ref.

0
[13]
99,800
[13]
39,717
[13]
870
[35]
0
[36]
14,070
[35]
8,235
[13]
5,148
[13]
3,418
[13]
9,500
[13]
11,935
[13]
16,800
[14]
14,700
[14]
1,450
This work
28,900
[14]
15,000
[14]
6,100
[14]
0
[14]
0
[37]
0
[14]
0
[14]
0
[14]
70,000
[38]
83,600
[38]
0
[13]
0
[13]
0
[39]
27,500
[14]
27,500
[14]
0
[40]
72,700
[13]
0
[13]
380
[17]
20,217
[17]
29,500
This work
42,920
[41]
38,950
[13]
447
[13]
4,928
[13]
5,148
[13]
3,080
[13]
5,773
[13]
5,773
[13]
4,400
[13]
29,400
[13]
41,400
[13]
2,583
[13]
8,235
[13]
5,148
[13]
9,500
[13]
4,400
[13]
0
[42]
13,522
[43]
3,795
[43]
(continued on next page)

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
14

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.14 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Table A (continued)

54
55
56
57
58
59
60
61
62
63

C6 H5 C2 H5 + H C6 H5 C2 H4 + H2
C6 H5 C2 H4 C6 H5 CHCH2 + H
C6 H5 CHCH2 C6 H6 + C2 H2
OC6 H4 CH3 + H HOC6 H4 CH3
OC6 H4 CH3 C6 H6 + H + CO
HOC6 H4 CH3 + OH OC6 H4 CH3 + H2 O
HOC6 H4 CH3 + H OC6 H4 CH3 + H2
HOC6 H4 CH3 + H C6 H5 CH3 + OH
HOC6 H4 CH3 + H C6 H5 OH + CH3
HOC6 H4 CH3 + C6 H5 CH2 OC6 H4 CH3 + C6 H5 CH3

Aa

na

Ea

Ref.

2.65E+02
3.16E+13
1.60E+11
2.50E+14
2.51E+11
6.00E+12
1.15E+14
2.21E+13
1.20E+13
1.05E+11

3.4
0
0
0
0
0
0
0
0
0

1,003
50,670
58,438
0
43,900
0
12,400
7,910
5,148
9,500

[43]
[40]
[40]
[13]
[13]
[13]
[13]
[13]
[13]
[13]

a k = AT n exp(E/RT ), A units: mol, cm, s; E units: cal/mol. Reverse rate constants are calculated through the equilibf
rium constant with associated thermodynamic data.

Appendix B
Table B
n-Heptane subset with rate constants

573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598

599
600

C7 H16 PC4 H9 + I C3 H7
C7 H16 + O2 C7 H15 -1 + HO2
C7 H16 + O2 C7 H15 -2 + HO2
C7 H16 + HO2 C7 H15 -1 + H2 O2
C7 H16 + HO2 C7 H15 -2 + H2 O2
C7 H16 + OH C7 H15 -1 + H2 O
C7 H16 + OH C7 H15 -2 + H2 O
C7 H15 -1 C7 H15 -2
C7 H15 -2 C7 H15 -1
C7 H15 -1 + O2 C7 H15 O2
C7 H15 O2 C7 H15 -1 + O2
C7 H15 -2 + O2 C7 H15 O2
C7 H15 O2 C7 H15 -2 + O2
C7 H15 O2 C7 H14 O2 H
C7 H14 O2 H + O2 O2 C7 H14 O2 H
O2 C7 H14 O2 H HO2 C7 H13 O2 H
HO2 C7 H13 O2 H OC7 H13 O2 H + OH
OC7 H13 O2 H OC7 H13 O + OH
OC7 H13 O CH2 O + C5 H11 + CO
C7 H15 -1 C5 H11 + C2 H4
C7 H15 -2 PC4 H9 + CH2 CHCH3
C5 H11 C2 H4 + I C3 H7
C7 H16 + H C7 H15 -1 + H2
C7 H16 + H C7 H15 -2 + H2
I C3 H7 + O2 CH2 CHCH3 + HO2
CH2 CHCH3 + H (+M) I C3 H7 (+M)

Low pressure limit: 0.16400E+55 11.1


Troe centering: 0.10000E+01 0.10000E14
H2 O enhanced by 5.000E+00
H2 enhanced by 2.000E+00
CO2 enhanced by 3.000E+00
CO enhanced by 2.000E+00
I C3 H7 + H C2 H5 + CH3
PC4 H9 (+M) C2 H5 + C2 H4 (+M)
Low pressure limit: 0.18970E+56 0.11910E+02
H2 O enhanced by 5.000E+00

Aa

na

Ea

Ref.

3.16E+16
6.00E+13
4.00E+13
5.00E+13
3.36E+13
1.05E+10
9.40E+07
2.00E+11
2.00E+11
2.50E+12
2.20E+15
2.50E+12
2.20E+15
2.00E+11
5.60E+12
2.00E+11
1.00E+09
8.40E+14
2.00E+13
2.50E+13
1.20E+13
7.97E+17
1.88E+05
2.60E+06
6.10E+20
5.70E+09
936.4
260 3000

0
0
0
0
0
1
1.6
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1.4
2.8
2.4
2.9
1.2

81,020
52,820
50,190
20,430
17,690
1,590
35
18,120
18,120
0
27,960
0
26,960
17,010
0
17,010
7,500
43,020
15,000
28,920
28,203
29,876
6,286
4,469
7,910
874

[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[19]
[13]
[44]

5.00E+13
1.06E+13
0.32263E5

0
0

0
27,828

[45]
[13]

(continued on next page)


Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.15 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
15

Table B (continued)
Aa
H2 enhanced by 2.000E+00
CO2 enhanced by 3.000E+00
CO enhanced by 2.000E+00
PC4 H9 C2 H5 + C2 H4

601

na

2.50E+13

Ea

Ref.

28,800

[13]

a k = AT n exp(E/RT ), A units: mol, cm, s; E units: cal/mol. Reverse rate constants (where applicable) are calculated
f

through the equilibrium constant with associated thermodynamic data.

Appendix C
Table C
Isooctane subset with rate constants

602
603
604
605
606
607
608
609
610
611
612
613
614

C8 H18 + O2 AC8 H17 + HO2


Reverse Arrhenius coefficients:
AC8 H17 + O2 AC8 H17 OO
Reverse Arrhenius coefficients:
AC8 H17 OO AC8 H16 OOHB
Reverse Arrhenius coefficients:
AC8 H16 OOHB + O2 AC8 H16 OOHBOO
Reverse Arrhenius coefficients:
AC8 H16 OOHBOO OC8 H15 OOH + OH
C8 H18 + OH AC8 H17 + H2 O
OC8 H15 OOH OC8 H15 O + OH
AC8 H17 + O2 JC8 H16 + HO2
Reverse Arrhenius coefficients:
OC8 H15 O + O2 C2 H3 + 2 CH2 O +H2 CCCH2 + CH3 + HO2
AC8 H17 IC4 H8 + CH2 CHCH3 + CH3
JC8 H16 IC4 H8 + CH2 CHC.H2 + CH3
IC4 H8 + O2 C2 H3 + C2 H4 + HO2
C8 H18 + HO2 AC8 H17 + H2 O2

Aa

na

Ea

Ref.

1.00E+16
1.00E+12
1.00E+12
2.51E+13
1.14E+11
1.00E+11
3.16E+11
2.51E+13
8.91E+10
3.00E+13
3.98E+15
3.16E+11
3.16E+11
2.45E+13
1.28E+12
1.92E+12
2.00E+14
3.02E+12

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

49,000
0
0
27,400
22,400
11,000
0
27,400
17,000
3,000
43,000
6,300
19,500
32,000
49,000
49,000
35,900
14,700

This work
[20]
[20]
[20]
[20]
[20]
[20]
[20]
[20]
This work
[20]
This work
[20]
[46]
[46]
This work
[46]
This work

a k = AT n exp(E/RT ), A units: mol, cm, s; E units: cal/mol. Reverse rate constant in reaction (614) is calculated through
f

the equilibrium constant with associated thermodynamic data.

Appendix D
Table D
Cross reactions and rate constants

615
616
617
618
619
620
621
622
623
624
625
626

C7 H15 -1 + C6 H5 CH3 C6 H5 CH2 + C7 H16


C7 H15 -2 + C6 H5 CH3 C6 H5 CH2 + C7 H16
C6 H5 + C7 H16 C7 H15 -1 + C6 H6
C6 H5 + C7 H16 C7 H15 -2 + C6 H6
C6 H5 CH2 OO + C7 H16 C6 H5 CH2 OOH + C7 H15 -1
C6 H5 CH2 OO + C7 H16 C6 H5 CH2 OOH + C7 H15 -2
C6 H5 CH2 OOH C6 H5 CH2 O + OH
AC8 H17 + C6 H5 CH3 C6 H5 CH2 + C8 H18
C6 H5 + C8 H18 AC8 H17 + C6 H6
C6 H5 CH2 OO + C8 H18 C6 H5 CH2 OOH + AC8 H17
C8 H18 + C7 H15 -1 C7 H16 + AC8 H17
C8 H18 + C7 H15 -2 C7 H16 + AC8 H17

Aa

na

Ea

Ref.

1.00E+11
1.00E+11
9.00E+11
2.00E+11
1.40E+06
2.33E+06
1.50E+16
1.00E+11
9.00E+11
1.40E+06
9.00E+11
9.00E+11

0
0
0
0
2.4
2.4
0
0
0
2.4
0
0

12,000
12,000
15,000
12,500
16,700
15,000
42,000
12,000
15,500
16,700
13,500
14,500

[3]
[3]
[3]
[3]
This work
This work
[14]
[3]
[3]
This work
[47]
[47]

a k = AT n exp(E/RT ), A units: mol, cm, s; E units: cal/mol. Reverse rate constants are calculated through the equilibf

rium constant with associated thermodynamic data.


Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF
16

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.16 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()

Appendix E
Table E
Revised rate coefficients for laminar flame speed simulations

215
216
298

299

301
627
628
629
630
631
632
633

CH2 CHCHC.H + O2 C.HCHCHO + CH2 O


CH2 CHCHC.H + O2 CH2 CHCCH + HO2
H2 CCCH + H (+M) = H2 CCCH2 (+M)
Low pressure limit:
H2 O enhanced by 5.000E+00
H2 enhanced by 2.000E+00
CO2 enhanced by 3.000E+00
CO enhanced by 2.000E+00
O2 enhanced by 2.000E+00
C2 H2 enhanced by 2.000E+00
H2 CCCH + H (+M) = H3 CCCH (+M)
Low pressure limit:
H2 O enhanced by 5.000E+00
H2 enhanced by 2.000E+00
CO2 enhanced by 3.000E+00
CO enhanced by 2.000E+00
O2 enhanced by 2.000E+00
C2 H2 enhanced by 2.000E+00
H2 CCCH + O = CH2 O + C2 H
IC4 H8 = CH2 CHC.H2 + CH3
Reverse Arrhenius coefficients:
C6 H5 C2 H5 + O2 = C6 H5 C2 H4 + HO2
C6 H5 C2 H4 + O2 = HO2 + C6 H5 CHCH2
C6 H5 C2 H4 + HO2 = OH + C6 H5 CHO + CH3
C6 H5 CHCH2 + O = C6 H5 + CH2 HCO
C6 H5 CHCH2 + OH = C6 H5 CH2 + CH2 O
C6 H5 CHCH2 + OH = C6 H5 CHO + CH3

Aa

na

Ea

Ref.

1.00E+11
1.00E+08
1.66E+15
0.376E+46

0.0
2.0
0.4
8.52

0
10,000
0
6,293

This work
This work
[13]

1.66E+15
0.878E+46

0.4
8.9

0
7,974

2.50E+14
1.92E+66
2.09E+59
1.40E+12
7.00E+11
2.00E+12
3.00E+08
6.00E+12
1.00E+13

0.0
14.2
13.2
0
0
0
1.4
0
0

0
128,100
29,530
34,000
17,000
0
1,000
0
0

[13]

This work
[48]
[48]
[14]
This work
This work
This work
This work
This work

a k = AT n exp(E/RT ), A units: mol, cm, s; E units: cal/mol. Reverse rate constants are calculated through the equilibf
rium constant with associated thermodynamic data.

References
[1] G. Kalghatgi, SAE Paper 2005-01-0239, 2005.
[2] W.J. Pitz, N.P. Cernansky, F.L. Dryer, F.N. Egolfopoulos, J.T. Farrel, D.G. Friend, H. Pitsch, SAE Paper
2007-01-0175, 2007.
[3] J.C.G. Andrae, P. Bjrnbom, R.F. Cracknell, G.T.
Kalghatgi, Combust. Flame 149 (2007) 224.
[4] J. Herzler, M. Fikri, K. Hitzbleck, R. Starke, C. Schulz,
P. Roth, G.T. Kalghatgi, Combust. Flame 149 (2007)
2531.
[5] D.F. Davidson, B.M. Gauthier, R.K. Hanson, Proc.
Combust. Inst. 30 (2005) 11751182.
[6] B.M. Gauthier, D.F. Davidson, R.K. Hanson, Combust.
Flame 139 (2004) 300311.
[7] G. Kalghatgi, P. Risberg, H.E. ngstrm, SAE Paper
2003-01-1816, 2003.
[8] G.T. Kalghatgi, R.A. Head, SAE Paper 2004-01-1969,
2004.
[9] J.C.G. Andrae, Fuel 87 (2008) 20132022.
[10] C.V. Naik, W.J. Pitz, C.K. Westbrook, M. Sjberg,
J.E. Dec, J. Orme, H.J. Curran, J.M. Simmie, SAE
Trans. 114 (2005) 13811387.

[11] M. Fikri, J. Herzler, R. Starke, C. Schulz, P. Roth, G.T.


Kalghatgi, Combust. Flame 152 (2008) 276281.
[12] M.U. Alzueta, P. Glarborg, K. Dam-Johansen, Int. J.
Chem. Kinet. 32 (2000) 498522.
[13] R. Sivaramakrishnan, R.S. Tranter, K. Brezinsky, Proc.
Combust. Inst. 30 (2005) 11651173.
[14] R. Bounaceur, I. Da Costa, R. Fournet, F. Billaud, F.
Battin-Leclerc, Int. J. Chem. Kinet. 37 (2005) 2549.
[15] P.Q.E. Clothier, D. Shen, H.O. Pritchard, Combust.
Flame 101 (1995) 383386.
[16] C. Ellis, M.S. Scott, R.W. Walker, Combust. Flame 132
(2003) 291304.
[17] F.F. Fenter, B. Noziere, F. Caralp, R. Lesclaux, Int. J.
Chem. Kinet. 26 (1994) 171189.
[18] R.P. Lindstedt, L.Q. Maurice, Combust. Sci. Technol. 120 (1996) 119167.
[19] N. Peters, G. Paczko, R. Seiser, K. Seshadri, Combust.
Flame 128 (2002) 3859.
[20] S. Tanaka, F. Ayala, J.C. Keck, Combust. Flame 133
(2003) 467481.
[21] J. Andrae, D. Johansson, P. Bjrnbom, P. Risberg, G.
Kalghatgi, Combust. Flame 140 (2005) 267286.

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

ARTICLE IN PRESS
JID:CNF

AID:7033 /FLA

CNF:7033

[mNY4a; v 1.91; Prn:13/06/2008; 14:05] P.17 (1-17)


J.C.G. Andrae et al. / Combustion and Flame ()
17

[22] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, J.A. Montgomery Jr.,
T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S.
Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi,
G. Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M.
Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa,
M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai,
M. Klene, X. Li, J.E. Knox, H.P. Hratchian, J.B. Cross,
C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann,
O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W.
Ochterski, P.Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich,
A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick,
A.D. Rabuck, K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J. Cioslowski,
B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I.
Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. AlLaham, C.Y. Peng, A. Nanayakkara, M. Challacombe,
P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C.
Gonzalez, J.A. Pople, Gaussian 03, Revision C.03,
Gaussian, Inc., Pittsburgh, PA, 2003.
[23] F. Neese, T. Schwabe, S. Grimme, J. Chem. Phys. 126
(2007) 124115124129.
[24] S. Grimme, J. Chem. Phys. 124 (2006) 3410834123.
[25] T. Schwabe, S. Grimme, Phys. Chem. Chem. Phys. 8
(2006) 43984401.
[26] J. Bengtsson, M.Sc. thesis, Royal Institute of Technology, Stockholm, 2006.
[27] R.J. Kee, F.M. Rupley, J.A. Miller, M.E. Coltrin, J.F.
Grcar, E. Meeks, H.K. Moffat, A.E. Lutz, G. DixonLewis, M.D. Smooke, J. Warnatz, G.H. Evans, R.S.
Larson, R.E. Mitchell, L.R. Petzold, W.C. Reynolds,
M. Caracotsios, W.E. Stewart, P. Glarborg, C. Wang,
C.L. McLellan, O. Adigun, W.G. Houf, C.P. Chou, S.F.
Miller, P. Ho, P.D. Young, D.J. Young, D.W. Hodgson,
M.V. Petrova, K.V. Puduppakkam, CHEMKIN Release
4.1, Reaction Design, San Diego, CA, 2006.
[28] D.F. Davidson, R.K. Hanson, Int. J. Chem. Kinet. 36
(2004) 510523.
[29] K. Fieweger, R. Blumenthal, G. Adomeit, Combust.
Flame 109 (1997) 599619.
[30] J. Herzler, L. Jerig, P. Roth, Proc. Combust. Inst. 30
(2005) 11471153.

[31] S. Tanaka, F. Ayala, J.C. Keck, J.B. Heywood, Combust. Flame 132 (2003) 219239.
[32] S.G. Davis, H. Wang, K. Brezinsky, C.K. Law, Proc.
Combust. Inst. 26 (1996) 10251033.
[33] Z. Zhao, J.P. Conley, A. Kazakov, F.L. Dryer, SAE Paper 2003-01-3265, 2003.
[34] D.R. Lide (Ed.), CRC Handbook of Chemistry and
Physics, 88th ed., 20072008.
[35] D.L. Baulch, C.J. Cobos, R.A. Cox, P. Frank, G. Hayman, Th. Just, J.A. Kerr, T. Murrells, M.J. Pilling, J.
Troe, R.W. Walker, J. Warnatz, J. Phys. Chem. Ref.
Data 23 (1994) 8471033.
[36] A. Hoffmann, M. Klatt, H.Gg. Wagner, Z. Phys.
Chem. 168 (1990) 112.
[37] J.A. Miller, C.F. Melius, Combust. Flame 91 (1992)
2139.
[38] M.B. Colket, D.J. Seary, Proc. Combust. Inst. 25
(1994) 883891.
[39] H. Hippler, C. Reihs, J. Troe, Proc. Combust. Inst. 23
(1991) 3743.
[40] W. Mller-Markgraf, J. Troe, J. Phys. Chem. 92 (1988)
49144922.
[41] K. Brezinsky, T.A. Litzinger, I. Glassman, Int. J. Chem.
Kinet. 16 (1984) 10531074.
[42] D.L. Baulch, C.J. Cobos, R.A. Cox, P. Frank, G. Hayman, Th. Just, J.A. Kerr, M.J. Pilling, J. Troe, R.W.
Walker, J. Warnatz, J. Phys. Chem. Ref. Data 21 (1992)
411734.
[43] D.L. Baulch, C.T. Bowman, C.J. Cobos, R.A. Cox, Th.
Just, J.A. Kerr, M.J. Pilling, D. Stocker, J. Troe, W.
Tsang, R.W. Walker, J. Warnatz, J. Phys. Chem. Ref.
Data 34 (2005) 7571397.
[44] P.W. Seakins, S.H. Robertson, M.J. Pilling, I.R. Slagle,
G.W. Gmurczyk, A. Bencsura, D. Gutman, W. Tsang,
J. Phys. Chem. 97 (1993) 44504458.
[45] W. Tsang, J. Phys. Chem. Ref. Data 17 (1988) 887
952.
[46] M. Jia, M. Xie, Fuel 85 (2006) 25932604.
[47] P.A. Glaude, V. Conraud, F. Fournet, F. Battin-Leclerc,
G.M. Come, G. Scacchi, P. Dagaut, M. Cathonnet, Energy Fuels 16 (2002) 11861195.
[48] H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook,
Combust. Flame 129 (2002) 253280.

Please cite this article in press as: J.C.G. Andrae et al., HCCI experiments with toluene reference fuels modeled by a
semidetailed chemical kinetic model, Combust. Flame (2008), doi:10.1016/j.combustflame.2008.05.010

You might also like