Anda di halaman 1dari 8

Available online at www.sciencedirect.

com

Proceedings
of the

Proceedings of the Combustion Institute 32 (2009) 28432850

Combustion
Institute
www.elsevier.com/locate/proci

Experimental and kinetic modeling study of the eect


of fuel composition in HCCI engines
Marco Mehl a,*, Tiziano Faravelli a,1, Eliseo Ranzi a,1, David Miller b,2,
Nicholas Cernansky b,2
b

a
CMIC Politecnico di Milano, P.zza Leonardo da Vinci, 32, 20133 Milano, Italy
Department of Mechanical Engineering and Mechanics, Drexel University, 3141 Chestnut Street,
Philadelphia, PA 19104, USA

Abstract
The chemical behavior of three dierent fuel surrogates, containing n-heptane, iso-octane, toluene, and
1-pentene, under HCCI conditions is analyzed. New experimental measurements of the in-cylinder composition have been carried out on a running HCCI engine equipped with a special customized cylinder head
allowing the direct sampling of the gas.
These data were compared with calculation obtained from a detailed kinetic mechanism coupled with a
1-D/quasi-D engine model. This model allows to study the autoignition and the combustion behavior of
dierent fuel surrogates. After a preliminary analysis on the model parameters and of stoichiometry eect,
the combustion chemistry of dierent fuel surrogates is discussed. Model calculations correctly reproduce
the combustion behavior of the three surrogates. Moreover the parametric analysis allowed better interpretation of the engine measurements. The coupled approach adopted in this work showed to be suited to analyze the eect of dierent fuel formulation and help the synergistic development of both engines and fuels.
Published by Elsevier Inc. on behalf of The Combustion Institute.
Keywords: Detailed kinetics; HCCI; Fuel surrogates; Composition measurements; Engine modeling

1. Introduction
While modern vehicles and energy conversion
processes are cleaner than ever, increasing levels
of industrialization and the growth of world population requires increasing diligence with respect
to environmental and energy issues. To address
*

Corresponding author. Fax: +39 02 7063 8173.


E-mail addresses: marco.mehl@polimi.it (M. Mehl),
tiziano.faravelli@polimi.it (T. Faravelli), eliseo.ranzi@
polimi.it (E. Ranzi), dmiller@coe.drexel.edu (D. Miller),
cernansk@coe.drexel.edu (N. Cernansky).
1
Fax: +39 02 7063 8173.
2
Fax: +1 215 895 1478.

these issues both European Union and U.S. institutions regularly propose stricter regulations
which require the use of more ecological technologies in both automotive and energy industries.
Traditional Spark Ignition (SI) and Compression Ignition (CI) engines are continuously
improved in order to reduce fuel consumption
without lessening the performance. Considerable
eort has also been invested to develop new operating paradigms such as HCCI (Homogeneous
Charge Compression Ignition) engines that oer
the prospect of combining the best features of
both SI and CI. This combustion strategy is characterized by high dilutions to reduce the temperature peaks and to limit NOx formation and by

1540-7489/$ - see front matter Published by Elsevier Inc. on behalf of The Combustion Institute.
doi:10.1016/j.proci.2008.08.001

2844

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

premixed, gas phase combustion to avoid soot


production. Combustion phasing is entirely controlled by chemical kinetics. Unfortunately HCCI
is best suited to steady-state, mid-range conditions
and hybrid systems which have the capability of
switching between HCCI combustion and either
SI or CI are needed for other conditions. Due to
the kinetic control of HCCI, developing the complex control system necessary to maintain stable
and exible combustion requires understanding
the in-cylinder chemistry [1].
Fuel composition has also been rapidly evolving. In recent years oil companies have started
oering gasolines with higher octane number, diesel fuels with high performance, and biofuels with
a range of formulations. These developments have
renewed interest in the fundamental study of
hydrocarbon oxidation to investigate and ultimately control the autoignition behavior of the
new generation fuels.
With the increasing importance of simulation
to engine design, the need for detailed kinetic
mechanisms of hydrocarbon oxidation has also
increased. Unfortunately, modern fuels are complicated mixtures of hundreds of hydrocarbons
which are not amenable to developing chemical
kinetic mechanisms directly. The modern technique to overcome this problem is to develop
mechanisms for simpler mixtures of selected
hydrocarbons whose oxidation mimics that of
the real fuel. These simpler mixtures have been
called surrogates [2].
In this work the chemical behavior of dierent
possible fuel surrogate components under HCCI
conditions is analyzed. Experimental measurements of the in-cylinder composition have been
carried out on a running HCCI engine and a
detailed kinetic mechanism, coupled with a 1-D/
quasi-D engine model, is used to study the chemical evolution of the fresh charge during combustion. After a preliminary parametric analysis on
the model parameters and the eect of stoichiometry, the combustion chemistry of dierent fuel
surrogates is discussed. The model allowed better
interpretation of the experimental measurements
and its use can be conveniently extended to the
study of dierent surrogate blends, in order to test
the fuel eect on engine performance.
2. Experimental apparatus and procedures
The experiments were carried out in a single
cylinder, four stroke, air cooled research engine
which was originally modied at Sandia National
Laboratory with further modications at Drexel
University [35]. In order to provide sampling
access from the centerline of the pancake combustion chamber, a customized cylinder head with
side-mounted valves and spark plugs (not red
in this study) was employed. In its present cong-

uration, the engine has a 76.2 mm bore, 82.6 mm


stroke and 8.2 compression ratio.
In this study, the inlet air temperature was
423 K and the inlet manifold pressure was set at
1 atm. The liquid fuel was introduced into the
heated inlet air stream by a commercial automotive liquid fuel injector located about 35 tube
diameters from the intake port. To insure complete mixing, the fuel air mixture is passed
through a 15 cm section of small metal beads followed by a 15 cm turbulator consisting of countercurrent vanes. This system has been proven to
produce a fully vaporized, homogeneous fuel air
mixture in previous studies [36]. The fuel delivery
system was capable of maintaining the equivalence ratio within 5% of the desired set point.
The engine was operated at a constant engine
speed of 750 rpm and inlet air ow rate of 65
Std l/min (volumetric eciency of 71%; light
load). These conditions established stable HCCI
operation (<3% CoV in IMEP) which allowed
in-cylinder sampling to generate species evolution
proles. At each experimental condition, a representative cylinder pressure trace was input to a
thermodynamic model for calculation of temperature and heat release proles. The main features of
this engine are summarized in Table 1. A more
detailed description of the device can be found
in previous publications [36]. Since the main
interest of this study involved the study of the
in-cylinder chemistry, no BMEP measurements
were systematically carried out.
In-cylinder gas samples were extracted by an
electromagnetic sampling valve located at the centerline of the combustion chamber where core gas
samples free of boundary layer eects can be
obtained [7]. The valve was programmed to open
for 2 ms (9 CAD at 750 rpm) at selected initial
crank angles, giving time/CAD integrated samples. To acquire sucient samples the valve was
opened on every engine cycle for a period of 60
to 100 cycles. The sampled gases were stored in
heated sample loops (held at 70 C) and pressurized to 170 kPa. The extracted samples were analyzed using two independent gas chromatographic
Table 1
Drexel engine main specications
Bore
Stroke
Displacement
C.R.
Inlet Pressure
Speed
Inlet Temperature
Fuel
Equivalence Ratio
Flow rate
Volumetric Eciency
IVC
EGR

0.0762 m
0.0826 m
377.69 cm
8.2:1
760 torr
750 rpm
423 K
PRF20
0.5
65 l/min
71%
235 CAD
10%

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

columns each with equipped with ame ionization


detectors (FID). For CO, CO2, formaldehyde,
propionaldehyde, acetone and other small oxygenates a series conguration of Porapak Q and
N packed columns were coupled with a ruthenium
methanizer to allow separation and measurement
on the FID. For the fuels and major hydrocarbon
species, DB-1 capillary columns provided the separation. Both column systems used temperature
programming of the column temperature to
improve separation. Further details of the gas
sampling and analysis system are available elsewhere [58].
3. Kinetic mechanism
The kinetic modeling of autoignition of hydrocarbon fuels in engines requires careful analysis of
both low and high-temperature mechanisms. A
general detailed kinetic scheme, consisting of
more than 300 species and 8000 reactions, has
been discussed in previous papers [9,10]. A brief
summary of the fundamentals of the approach
will be presented here below.
The kinetic scheme is based on a hierarchical
modularity. The core of the model is a detailed
sub-mechanisms of H2 and C1C4 species, while
the modular structure allows the introduction of
dierent reaction subsystems required to describe
the oxidation of higher hydrocarbons or the formation of specic pollutants.
Assuming rules for analogous reactions, only a
few main classes of reactions and their fundamental kinetic parameters are needed to extend the
scheme to heavier species. Automatic procedures
allow the generation of detailed mechanisms for
long chain hydrocarbons. The number of species
involved in these sub-mechanisms increase exponentially with the number of carbon atoms.
Lumping procedures, as described previously
[11], are utilized to reduce the dimensionality of
these schemes.
The mechanism presented here has been validated in a wide range of operating conditions
through the comparison with experimental measurements carried out in well controlled reaction
environments (jet stirred reactors, rapid compression machines, shock tubes, ow reactors,. . .)
[9,10]. The comprehensive validation proved the
reliability of the mechanism in all the conditions
relevant to engine combustion, extending the possible elds of investigation far beyond the conditions discussed here. The kinetic scheme, which
is capable of simulating the combustion behavior
of real transportation fuels [12], also has been successfully applied to the evaluation of the autoignition propensities of dierent fuel surrogate
components and mixtures in SI and HCCI engine
over a wide set of operating condition, as discussed in previous works [1317]. A recent devel-

2845

opment of the mechanism included the lowtemperature reactions of unsaturated hydrocarbons. Nowadays alkenes are commonly added in
fuel formulation as octane improvers. Like their
saturated homologues, the mechanism involves
the initial formation of alkenyl radicals and their
successive oxidations with the possible degenerate
branching path with peroxide and ketohydroperoxide species at low-temperature conditions.
Two major dierences from the alkane oxidation mechanism must be highlighted. The rst is
the formation of very stable allylic radicals and
the second is the addition of propagating radicals
to the double bond. One of the reasons for the
lower reactivity of alkenes, when compared to
alkanes, is the formation of allyl radicals. As a
consequence of their relative stability, these radicals reach high concentrations and are available
for oxygen addition and/or recombination reactions. Moreover, the O2 addition reactions need
to overcome this stability, while the corresponding
allyl-peroxy radicals decompose in a favored way.
On these bases, the ceiling temperature shifts at
lower temperatures with the result of a limited
importance of the low-temperature mechanism.
The second alkene peculiarity relates to radical
addition to the double bond. These reactions
mainly involve H, OH or HO2 radicals, respectively leading to the formation of alkyl radicals,
hydroxyalkyl or hydroperoxyalkyl radicals, which
further contribute to the low temperature reactivity. In agreement with previous kinetic analysis
[18,19], below 650 K the foremost reaction pathway involves the formation of hydroxyalkyl radicals and the successive oxygen addition. Due the
high electronegativity of oxygen, the net result
of this O2 addition is the nal decomposition to
form OH radical and two aldehydes, via the
Waddington mechanism. The successive reactions
of alkyl and hydroperoxyalkyl radicals obtained
via H and HO2 additions are already discussed
and considered in the oxidation mechanism of
alkanes. Radical addition reactions prevail on
the corresponding H-abstraction reactions mainly
for small alkenes, such as ethylene and propylene.
The limited reactivity of propylene, when compared to heavier alkenes, is mainly due to the difculty of isomerizing the allyl peroxy radical
through the internal abstraction of the vinyl Hatoms. H abstraction reactions, not only limited
to the allyl H-atoms, prevail over the corresponding radical addition reactions above pentene [20].
Aromatics are a further family of hydrocarbons used to improve the octane rating of fuels.
The most relevant aromatic is toluene, which has
become a third component in a class of standard
SI reference fuels termed TRFs (Toluene Reference Fuels [21]), along with n-heptane and
iso-octane. Toluene reduces low-temperature
reactivity and, generally speaking, it can be considered as an antiknock component in fuel blends.

2846

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

The low-temperature reactivity of TRFs can be


almost totally assigned to the alkane fraction. At
low-temperature the most relevant pathways of
consumption of toluene are the H-abstraction
reactions, mainly performed by OH radicals.
Two out of four possible isomers of the radical
C7H7have been introduced in the kinetic model:
the rst one is the resonantly stabilized benzyl
radical C6H5CH2, while the second one,
C6H4CH3, is a non- resonantly pseudospecies,
representing the three aryl radicals, all of which
behave in a similar way. As expected, the benzyl
radical has a much higher concentration than
C6H4CH3, due to its higher stability. In the lowtemperature region benzyl radical recombines
with HO2 radical forming an unstable peroxide
which successively decomposes to the benzyl alcohol radical releasing OH.
OH radical can also substitute the methyl
group, leading to the formation of phenol and a
slow reacting methyl radical. The conversion from
OH to CH3 depresses the general reactivity. This
reaction, which is not so relevant when pure toluene is considered, becomes important in mixtures
where other components act as OH sources. A
further reaction pathway is the O addition to toluene which leads to the formation of cresoxy
radical.
At higher temperatures the benzyl radical also
reacts with O and H leading to the formation of
benzaldehyde (rapidly consumed through an Habstraction and successive CO elimination) and
phenyl radical, respectively.
C6H4CH3 radical reacts with molecular oxygen
to form the cresoxy radical. This radical, analogously to the consumption of phenoxy C6H5O
(important to explain the benzene oxidation),
plays an important role in the toluene consumption with fulvene and benzene formation at high
temperatures. The successive oxidation of benzene, through CO elimination and cyclo-pentadienyl formation, is included in the kinetic model
with the established literature parameters [22].
The kinetic model also includes a module
describing NOx formation mechanisms and their
specic interactions with the hydrocarbon oxidation. NO has a promoting eect on the low-temperature oxidation due to its direct interaction
with ROO radicals as follows:
ROO NO ! RO NO2

where R can be both an H atom or an alkyl group.


If R is hydrogen, the consumption of the relatively
slowly reacting radical HO2 to yield aggressive
OH radicals increases the reactivity, otherwise,
RO radicals decompose to small reactive alkyl
radicals and an aldehyde. A complete description
of the reactions involved in this process can be
found in [23]. A downloadable version of the
mechanism in Chemkin format can be found at:
http://www.chem.polimi.it/CRECKModeling/.

4. Engine simulation
A unique 1-D engine simulation code developed
at Politecnico di Milano [2426] was used in this
study. It includes eective and robust numerical
methods widely tested on chemical integration
problems [27], which overcome the diculties arising from the steep gradients typical of the combustion inside cylinders. This tool is regularly applied
to the study and optimization of SI, Diesel and
HCCI engines. In this study a multizone model
modied to account for HCCI operation is used
to analyze the combustion of dierent fuel surrogate components. The model, based on a 1-D
approach, simulates the whole engine system
accounting for all of the typical boundaries encountered in the ow systems. The present single-cylinder test engine was modeled accounting for the
intake and exhaust systems geometry, valve lift proles and timing to provide the correct boundary
conditions of temperature, pressure, internal
EGR and chemical compositions.
The simulation of HCCI combustion is
achieved by means of a multi-zone approach
which includes the previously discussed detailed
kinetic mechanism to describe the low-temperature reactions of the fresh mixture during compression and combustion. When the valves are
closed, the cylinder is divided into several zones,
accounting for temperature and composition
stratications. Assuming a uniform pressure over
the whole combustion chamber the temperature
of each zone (Ti), the pressure (p) and the chemical composition of each zone (xi, j) is calculated
according the following equations:
P

d
j xi;j ui;j mi
dQi
dmi vi dmi

p
hi
2
dh
dh
dhh
dh i
 1 dV  PNzone
 i


i1 xi m RpC dQ
 H_ reac
dh
m dh
dp
i
p;i
i

3
PNzone h v2i
vi
dh
i1 xi C  p
p;i

dxi;j
x_ i;j
dh

xi, j refers to the mass fraction of a chemical species j within the zone i, ui to the specic internal
energy and mi to the total mass of the zone. The
term H_ reac represents the sum of the heat released
by chemical reactions in each zone; Qi is the heat
exchange and is calculated by a revised version of
the Woschni model [25].
The model assumes a boundary layer, involved
in heat transfer, and an adiabatic core which is
divided into zones according to a temperature distribution centered on the mean temperature value
estimated by the 1-D model. Since the cylinder
modeling is achieved by a 0-D approach, thermal
stratications are accounted for without analyzing
their actual locations but assuming a statistical
distribution of conditions. The variance of the dis-

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

The stronger eect on the start of the high-temperature combustion due to the air/fuel ratio variation depends on the temperature rise produced
by the low-temperature heat release. The 5% variation in air/fuel ratio implies a deviation of about
10 K at 10 CAD BTDC. Such dierences
explain why a leaner or richer composition determines the main ignition timing which occurs when
the critical condition (950 K for the selected fuel)
is achieved.
These eects are consistent with composition
proles. The fuel consumption rate signicantly
increases when U rises, while smaller dierences
are obtained changing the distribution range. Differences in the fresh charge composition obviously
aect also the intermediate proles: richer charges
result in higher and narrower peak shapes as will
be discussed later.

tribution is extrapolated from CFD studies performed by Aceves et al. [28]. The role of thermal
and composition inhomogeneities is discussed in
a previous work [17], where a more detailed
description and the validation of the approach
can be found.
5. Results
The uid-dynamic model of the described
engine was previously discussed and validated
on the basis of experimental pressure, temperature
and species prole comparisons [17]. In this previous study the model, calibrated on a pressure
curve measured at U = 0.4, correctly predicted
the eect of the equivalence ratio, providing satisfactory species proles at dierent stoichiometries.
The sensitivity analysis on the eect of the temperature distribution assumed in the model at
valve closure is particularly important because
the variance of the temperature distribution is
very sensitive to uncertainties and it cannot be
directly evaluated. Figure 1 shows the simulated
temperature prole of the reference case selected
(U = 0.5, PRF20). The temperature stratication
assumed is based on a normal distribution having
a standard deviation equal to 20 K (r = 20 K).
The temperature trace (Fig. 1a) is compared with
the calculated curves obtained considering a different initial temperature dispersion (assuming
respectively r = 15 K and r = 25 K). A second
sensitivity analysis was undertaken to assess the
eect of 5% deviations from the desired stoichiometry due to the fuel delivery system precision.
Figure 1b shows the temperature proles obtained
at U = 0.475, U = 0.500 and U = 0.525.
It is evident that the stoichiometry variations
are much more inuential on the main heat release
timing than the temperature prole variations.
Most of the heat release is associated with the
ignition of the zones closer to the mean temperature value. Consequently wider distribution tails
do not aect the general combustion behavior.
Small dierences of the low-temperature oxidation onset can be detected.

6. Fuel eect
In the following paragraph the eect of dierent fuel composition is discussed on the basis of
new experimental measurements and computational results.
Three dierent surrogates were considered: a
traditional PRF20, a PRF20/1-pentene mixture
(90/10 vv), and a PRF20/toluene mixture (90/10
vv). The air/fuel ratio is xed at 0.5 in all the simulations. Calculated species proles are compared
with measurements paying attention to the formation of the combustion intermediate evolutions.
Figure 2 refers to the in-cylinder mean temperature proles obtained for the three mixtures. The
addition of neither 1-pentene nor toluene aects
the low-temperature (LT) reaction on-set but
delays the main combustion phasing by a couple
of CAD. The main reason for this behavior is that
the most reactive component of the fuel (n-heptane) controls the low-temperature reactivity.
Both toluene and 1-pentene are scarcely reactive at these conditions: at about 6 atm and
735 K the induction time of 1-pentene (28 ms) is
more than two times greater than for n-heptane
(about 10 ms) and toluene takes several second
2100

2100

= 25K

=0.525

1850

1800

=0.500

T [K]

1600

=0.475

1500

= 20K

1350

1200

1100

= 15K

900
600
680

2847

700

720
CAD

850

740

600
680

700

720

740

CAD

Fig. 1. Eect of the standard deviation and of the air/fuel ratio on temperature proles Fuel eect on temperature
proles.

2848

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850


2100

PRF20 + 10% pentene

T [K]

1800

1500

PRF20

1200

900

PRF20 + 10% toluene


600
680

700

720

740

CAD

Fig. 2. Fuel eect on temperature prole.

to ignite. n-Heptane drives the reactivity of the


system because its low-temperature branching
routes become active in the last part of the compression stroke. While 1-pentene shows a lowtemperature reactivity, the presence of the double
bond inhibits the propagation of that mechanism.
As a consequence the pentenyl radicals formed
during the fuel oxidation preferably lead to alternative decomposition pathways. When the lowtemperature mechanism starts, about 735 K, more
than the 60% of the reacting n-heptane is converted to ketohydroperoxydes while only 40% of
1-pentene is converted. This ratio becomes even
more relevant when temperature increases: at
about 825 K and 8 bar (mean cylinder conditions), when the low-temperature reactions are
nearly extinguishing, n-heptane is almost twice
as eective as 1-pentene in producing branching
intermediates (33% vs. 18%). Toluene does not
show low-temperature reactivity and its consumption is almost negligible during the LT phase.
The transition to the high-temperature (HT)
oxidation mechanism requires the in-cylinder temperature to exceed 950 K. The minimal reactivity
of the two additives considered reduces the rst
heat release associated with LT reactions and
the main combustion is then delayed.
Further insight into the fuel composition eect
is obtained observing the species proles. A huge
number of oxidation intermediates were measured
(n-heptane, iso-octane, CO, CO2, alkenes from C2
up to C7,methane, formaldehyde, acetaldehyde,
propionaldehyde, butylaldehyde, trimethylaldehyde, methanol, acetone, propylene oxide, methyl
vinyl ketone), only few of them are here reported
because of lack of space.
Figure 3 shows the comparison between experimental measurements of the most representative
compounds with calculations, i.e., the concentration proles of n-heptane, carbon monoxide, 1pentene, methane and acetaldehyde which respectively represent the fuel consumption, unsaturated, saturated, and oxygenated primary
products.

Since the measured concentration values were


obtained by collecting the in-cylinder gases over
a 2 ms interval, the experimental proles lead
the reaction timing. This eect was discussed in
a previous work where the experimental temperature and pressure proles were compared with
composition measurements [17]. As a matter of
fact, the measured concentration of each species
is an integral over several crank angle degrees.
To compensate for this eect, the predicted species
proles are shifted by about 9 CAD, in agreement
with the sampling duration at the considered
engine speed.
The model reproduces with a satisfactory accuracy the species proles measured for the three
dierent surrogates. The qualitative trend is fully
reproduced as well as the concentration values.
The most relevant discrepancy is detected for the
acetaldehyde peak, where the model slightly
underpredicts the experimental data relative to
the pure PRF. This deviation is consistent with
the measured amount of n-heptane detected at
valve closure at the same condition which is also
underestimated. Figure 4 shows the simulation
results obtained at U = 0.475 and U = 0.525 (corresponding to the experimental uncertainties)
compared with experimental data.
The measured value is within the calculated
range, in particular the measured intermediate
peak is very close to the upper limit. This result
is consistent with other species comparisons where
the peak values of the intermediates are underestimated in the pure PRF calculations.
The 1-pentene prole requires further discussions. Since 1-pentene is at the same time a reactant and a product during the low-temperature
oxidation of n-heptane, both experimental and
calculated proles show that 1-pentene concentration decreases signicantly only during the hightemperature phase. Consequently, even though
its net ux is almost negligible, 1-pentene contributes to the formation of oxygenated products and
branching intermediates.
7. Conclusion
The work presented a new set of experimental
data on the chemical evolution of three fuel surrogates in a HCCI engine. These experimental data
were compared with the predictions obtained by a
1-D/quasi-D uid dynamic engine model coupled
with a detailed kinetic mechanism.
Calculations allowed evaluation of the eect of
the deviations of the air/fuel ratio from the target
value. Small dierences in the air/fuel ratio signicantly inuence the total ignition timing, while
the low temperature step is scarcely aected. On
the contrary, the eect of model parameters (e.g.,
temperature distribution proles) resulted to be of
less importance in the considered condition.

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

8000

2849

1800

n-heptane
6000

CH3CHO

ppm

1200

4000

CO
600

2000

0
650

1-pentene
CH4

CH2O
660

670

680

690

700

710

720

b 8000

650

660

670

680

690

700

710 720

1600

CH3CHO

n-heptane
6000

1200

ppm

1-pentene
4000

800

CO

2000

0
650

400

CH4

CH2O
660

670

680

690

700

710

720

c 8000

650

660

670

680

690

700

710

720

1400
1200

n-heptane

ppm

6000

1000

CH3CHO

800

4000

CO

600
400

2000

1-pentene

200

0
650

CH2O
660

670

680

690

700

CH4

710

720

650

660

670

680

CAD

690

700

710

720

CAD

Fig. 3. Experimental and calculated species proles during combustion: (a) PRF20; (b) PRF20 + 10% 1-pentene;
(c) PRF20 + 10% toluene.

The model was able to reproduce the main features of the combustion as well as the eect of the
fuel, correctly predicting the measured species
1800

=0.500

=0.525

ppm

1200

=0.475

600

proles in the combustion chamber. The comparison with experimental measurements further conrms that the adopted approach is reliable in
investigating HCCI combustion. Although a limited set of conditions has been discussed here,
the exibility of the uid dynamic model together
with the generality of the detailed kinetic model,
which accounts for the oxidation of hydrocarbons
up to 16 C atoms and with the pollutant formation mechanisms (NOx, PAHs and soot), make
this tool suited to the study of dierent fuel formulation in the context of the synergistic development of both engines and fuels.

Acknowledgments

0
650

660

670

680

690

700

710

720

CAD

Fig. 4. Eect of the air/fuel ratio on the acetaldehyde


peak. (PRF20).

The authors of the CMIC Department of


Politecnico di Milano acknowledge the nancial
support of Eni.

2850

M. Mehl et al. / Proceedings of the Combustion Institute 32 (2009) 28432850

Appendix A. Supplementary data


Supplementary data associated with this article
can be found, in the online version, at
doi:10.1016/j.proci.2008.08.001.
References
[1] H. Ando, Future Powerplants, 10th International
Conference on Present and Future Engines for
Automobiles, Rhodes, May 2731, 2007.
[2] W.J. Pitz, N.P. Cernansky, F.L. Dryer, F.N. Egolfopoulos, J.T. Farrell, D.G. Friend et al., Development
of an Experimental Database and Chemical Kinetic
Models for Surrogate Gasoline Fuels, SAE Paper No.
2007-01-0175, SAE Trans., 2007.
[3] J. Zheng, W. Yang, D.L. Miller, N.P. Cernansky,
Prediction of Pre-ignition Reactivity and Ignition
Delay for Using a Reduced Chemical Kinetic Model,
SAE Paper No. 2001-01-1025, SAE Trans., 2001.
[4] Y. Henig, S. Addagarla, D.L. Miller, R.D. Wilk,
N.P. Cernansky, Autoignition of n-Butane/Isobutane Blends in a Knock Research Engine, SAE
Paper No. 890157, SAE Trans., 96, 1989.
[5] W. Yang, J. Zheng, D.L. Miller, N.P. Cernansky, Tracer
Fuel Injection Studies on Exhaust Port Hydrocarbon
Oxidation: Part II, SAE Paper No. 2000-01-1945, 2000.
[6] W. Yang, The Chemistry Controlling Post Combustion Hydrocarbon Oxidation and Homogeneous
Charge Compression Ignition, Ph.D. Dissertation,
Drexel University, Philadelphia, PA, 2002.
[7] N.P. Cernansky, R.M. Green, W.J. Pitz, C.K.
Westbrook, Combust. Sci. Technol. 50 (1986) 325.
[8] R.M. Green, C.D. Parker, W.J. Pitz, C.K. Westbrook, The Autoignition of Isobutane in a Knocking
Research Engine, SAE Paper No. 870169, SAE
Trans., 1987.
[9] E. Ranzi, P. Gauri, T. Faravelli, P. Dagaut,
Combust. Flame 103 (1995) 91106.
[10] E. Ranzi, T. Faravelli, P. Gauri, A. Sogaro, A.
DAnna, A. Ciajolo, Combust. Flame 108 (1997)
2442.
[11] E. Ranzi, N. Dente, A. Goldaniga, G. Bozzano, T.
Faravelli, Prog. Energy Combust. Sci. 27 (2001) 99.
[12] E. Ranzi, Energy Fuels 20 (2006) 10241032.
[13] M. Mehl, T. Faravelli, E. Ranzi et al., Kinetic
Modeling Study of Octane Number and Sensitivity of

[14]
[15]
[16]

[17]

[18]
[19]
[20]
[21]
[22]
[23]
[24]

[25]

[26]
[27]
[28]

Hydrocarbon Mixtures in CFR Engines, SAE Paper


No. 2005-24-077, 2005.
M. Mehl, T. Faravelli, E. Ranzi et al., Kinetic
Modeling of Knock Properties in Internal Combustion Engines, SAE Paper No. 2006-01-3239, 2006.
M. Mehl, T. Faravelli, F. Giavazzi, et al., Energy
Fuels 20 (6) (2006) 23912398.
G. DErrico, T. Lucchini, A. Onorati et al.,
Development and Experimental Validation of a
Combustion Model with Detailed Chemistry for
Knock Predictions, SAE Paper No. 2007-01-0938,
2007.
M. Mehl, A. Tardani, T. Faravelli et al., A
Multizone Approach to the Detailed Kinetic Modeling of HCCI Combustion, SAE Paper No. 2007-240086, 2007.
S. Touchard, R. Fournet, P.A. Glaude, et al., Proc.
Combust. Inst. 30 (2005) 10731081.
R. Minetti, A. Roubaud, E. Therssen, M. Ribaucour, L.R. Sochet, Combust. Flame 118 (1999) 213
220.
S.K. Prabhu, R.K. Bhat, D.L. Miller, N.P. Cernansky, Combust. Flame 104 (1996) 377.
J.C.G. Andrae, P. Bjornbom, R.F. Cracknell, G.T.
Kalghatgi, Combust. Flame 149 (2007) 224.
K. Roy, C. Horn, P. Frank, V.G. Slutsky, T. Just,
Proc. Combust. Inst. 27 (1998) 329.
A. Frassoldati, T. Faravelli, E. Ranzi, Combust.
Flame 132 (12) (2003) 188207.
A. Onorati, G. DErrico, G. Ferrari, 1D Fluid
Dynamic Modeling of Unsteady Reacting Flows in
the Exhaust System with Catalytic Converter for S.I.
Engines, SAE Paper No. 2000-01-0210, SAE Trans.,
2001.
G. DErrico, G. Ferrari, A. Onorati, T. Cerri,
Modeling the Pollutant Emissions from a S.I.
Engine, SAE Paper No. 2002-01-0006, SAE
Trans., 2002.
G. DErrico, T. Cerri, T. Lucchini, Development and
Application of S.I. Combustion Models for Emissions
Prediction, SAE Paper No. 2006-01-1108, 2006.
G. Buzzi-Ferraris, D. Manca, BzzOde: Comput.
Chem. Eng. 22 (11) (1998) 15951621.
S.M. Aceves, D.L. Flowers, J. Martinez-Frias, F.
Espinosa-Loza, M. Christensen, B. Johansson
et al., Analysis of the Eect of Geometry Generated Turbulence on HCCI Combustion by MultiZone Modeling, SAE Paper No. 2005-01-2134,
2005.

Anda mungkin juga menyukai