Anda di halaman 1dari 16

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/228846056

Spark Ignition Engine Combustion Modeling


Using a Level Set Method with Detailed
Chemistry

Article · May 2006


DOI: 10.4271/2006-01-0243

CITATIONS READS

24 249

2 authors, including:

Rolf Reitz
University of Wisconsin–Madison
363 PUBLICATIONS 7,492 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CFD Modeling of Engine Flow and Combustion Processes View project

Combustion theory View project

All content following this page was uploaded by Rolf Reitz on 20 May 2014.

The user has requested enhancement of the downloaded file.


SAE TECHNICAL
PAPER SERIES 2006-01-0243

Spark Ignition Engine Combustion Modeling


Using a Level Set Method with
Detailed Chemistry
Long Liang and Rolf D. Reitz
University of Wisconsin – Madison

Reprinted From: Multi-Dimensional Engine Modeling 2006


(SP-2011)

2006 SAE World Congress


Detroit, Michigan
April 3-6, 2006

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760 Web: www.sae.org
The Engineering Meetings Board has approved this paper for publication. It has successfully completed
SAE's peer review process under the supervision of the session organizer. This process requires a
minimum of three (3) reviews by industry experts.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise,
without the prior written permission of SAE.

For permission and licensing requests contact:

SAE Permissions
400 Commonwealth Drive
Warrendale, PA 15096-0001-USA
Email: permissions@sae.org
Tel: 724-772-4028
Fax: 724-776-3036

For multiple print copies contact:

SAE Customer Service


Tel: 877-606-7323 (inside USA and Canada)
Tel: 724-776-4970 (outside USA)
Fax: 724-776-0790
Email: CustomerService@sae.org

ISSN 0148-7191
Copyright © 2006 SAE International
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE.
The author is solely responsible for the content of the paper. A process is available by which discussions
will be printed with the paper if it is published in SAE Transactions.

Persons wishing to submit papers to be considered for presentation or publication by SAE should send the
manuscript or a 300 word abstract to Secretary, Engineering Meetings Board, SAE.

Printed in USA
2006-01-0243

Spark Ignition Engine Combustion Modeling Using a Level Set


Method with Detailed Chemistry
Long Liang and Rolf D. Reitz
University of Wisconsin-Madison

Copyright © 2006 SAE International

ABSTRACT levels. There have been several approaches to model


the premixed and partially premixed combustion
A level set method (G-equation)-based combustion occurring in SI engines. They can be classified into
model incorporating detailed chemical kinetics has been turbulent mixing controlled, flamelet and PDF
developed and implemented in KIVA-3V for Spark- approaches. Abraham et al. [1] proposed a characteristic
Ignition (SI) engine simulations for better predictions of timescale combustion (CTC) model which used k-ε
fuel oxidation and pollutant formation. Detailed fuel model for turbulent transport and a combination of a
oxidation mechanisms coupled with a reduced NOX mixing-controlled and an Arrhenius-controlled species
mechanism are used to describe the chemical conversion rate. Flamelet models are another group of
processes. The flame front in the spark kernel stage is widely used methods which are either based on the
tracked using the Discrete Particle Ignition Kernel (DPIK) progress variable, c, or on the non-reacting scalar, G [2].
model. In the G-equation model, it is assumed that after The Bray-Moss-Libby (BML) model and the Coherent
the flame front has passed, the mixture within the mean Flame Model (CFM) are based on the progress variable,
flame brush tends to local equilibrium. The subgrid-scale c, which is viewed either as a normalized temperature or
burnt/unburnt volumes of the flame containing cells are as a normalized product mass fraction [3]. As an
tracked for the primary heat release calculation. A example, Boudier et al. [4] implemented an extended
progress variable concept is introduced into the turbulent version of CFM into KIVA to simulate turbulent flame
flame speed correlation to account for the laminar to ignition and propagation in SI engines, with focus on the
turbulent evolution of the spark kernel flame. To test the evolution criterion from the laminar stage to the fully
model, a homogeneous charge propane SI engine was developed turbulent stage based on flame stretch
modeled using a 100-species, 539-reaction propane effects. The model was validated by comparison with in-
mechanism, coupled with a reduced 9-reaction NOx cylinder flame front contours in an experimental SI
mechanism for the chemistry calculations. Good engine. The level set method (G-equation) is a powerful
agreement with experimental cylinder pressures and tool for describing interface evolution. With its
NOx data was obtained as a function of spark timing, application to combustion, Williams [5] first suggested a
engine speed and EGR levels. The model was also transport equation of a non-reactive scalar, G, for
applied to a stratified charge two-stroke gasoline engine laminar flame propagation. Peters [2,6] subsequently
simulations, and good agreement with measured data extended this approach to the turbulent flame regime.
was obtained. The turbulent G-equation concept has been successfully
applied to SI engine combustion simulations by Dekena
INTRODUCTION et al. [7], Tan [8] and Ewald et al. [9].

Multidimensional Computational Fluid Dynamics (CFD) In recent years, to better understand the fundamental
modeling has become an indispensable tool in the engine combustion process and to further improve the
design and analysis of low-emission, high-fuel-efficiency versatility of multidimensional models, attention is being
Internal Combustion Engines (ICE). Good understanding given to models incorporating comprehensive
of the in-cylinder turbulent combustion is one of the key elementary chemical kinetic mechanisms. A large
factors for successful modeling. In this paper, we focus amount of work has been done on developing detailed
on the combustion modeling of homogeneous charge chemical kinetic mechanisms for fuel oxidation and
and stratified charge SI engines. pollutant formation [10]. Further, research on
mechanism reduction and parallel computing techniques
The in-cylinder turbulent combustion in SI engines is a have made it computationally affordable to incorporate
complicated aero-thermo-chemical process especially the reduced detailed chemical kinetic mechanisms into
due to the turbulence and chemistry interactions on multidimensional engine simulations. The objective of
tremendously different time-scale and length-scale the current work is to incorporate detailed chemical
kinetics into the G-equation-based turbulent combustion 12
a4 b32 l ⎡ ⎛ a 4 b32 l ⎞ ⎤
2

model which was implemented into the KIVA-3V code by ST0 2 u ′l


0
= 1 − + ⎢ ⎜ ⎟ + a 4 b3 0 ⎥ (3)
Tan et al. [11,12]. Specifically, for SI engine combustion, SL 2b1 l F ⎢ ⎝ 2b1 l F ⎠ sL lF ⎥
⎣ ⎦
detailed fuel oxidation mechanisms coupled with a
reduced NOX mechanism are applied behind the mean G
flame front for modeling post flame combustion and NOX where vjf is the fluid velocity vector, Dt is the turbulent
formation. The chemical kinetic mechanisms are also diffusivity, and a4 , b1 , b3 , and cs are modeling constants
applied in front of the flame front for the potential from the turbulence model or experiment (cf. Peters [2]).
capability of predicting the compression autoignition of
k and ε are the Favre mean turbulence kinetic energy
the end-gas. Also, In the course of coupling detailed
chemistry with the G-equation combustion model for the and its dissipation rate from the RNG k - ε model [13].
primary heat release calculation within the flame front, u ′ is the turbulence intensity. SL0 is the unstretched
the laminar and turbulent flame speed correlations were laminar flame speed, and l and lF are the turbulence
required to be revisited and improved for a better integral length scale and laminar flame thickness,
description of the turbulent flame propagation process.
respectively. κi is the mean flame front curvature:
THE MODELS
⎛ ∇G i ⎞
κi = ∇⋅ ⎜ (4)
i ⎟
COMBUSTION MODEL WITH DETAILED CHEMISTRY ⎝ | ∇G | ⎠

G-equation description of turbulent flame propagation The flame thickness, lF , can be estimated by [14]:
In the flamelet modeling theory of premixed turbulent
combustion by Peters [2], two regimes of practical ( λ / c p ) |T0
lF = (5)
interest were addressed: the corrugated flamelet regime ρ u S L0
where the entire reactive-diffusive flame structure is
assumed to be embedded within eddies of the size of
where the λ / cp term is approximated by [15]:
the Kolmogorov length scale η ; and the thin reaction
zone regime where the Kolmogorov eddies can 0.7
penetrate into the chemically inert preheat zone of the g ⎛ T ⎞
λ / c p = 2.58 × 10−4 ⎜ ⎟ (6)
reactive-diffusive flame structure, but cannot enter the cm − sec ⎝ 298K ⎠
inner layer where the chemical reactions occur. Peters
derived level set equations applicable to both regimes, In this work the inner layer temperature, T0 , was
including the Favre averaged equations for the mean,
k
i , and its variance, G approximated as 1500K for the propane and iso-octane
G ′′2 , and a model equation for the flames considered.
flame surface area ratio, which in turn, gives an
algebraic solution for the steady-state planar turbulent In the implementations both by Tan [11,12] and Ewald
flame speed, ST0 . These equations together with the [9], a normal distance constraint | ∇G i |= 1 is applied
Reynolds averaged Navier-Stokes equations and the beyond the flame front surface. In this case, the
turbulence modeling equations, form a complete set to turbulent flame brush thickness, lF ,t , can be defined in
describe premixed turbulent flame front propagation [2].
Considering the Arbitrary Lagrangian-Eulerian (ALE) k
terms of G ′′2 as [2]:
numerical method used in the KIVA code, Tan [12]
modified the convection term of the G i transport k G
lF ,t = (G′′2 (x, t))1/ 2 |Gi =G (7)
equation to account for the velocity of the moving vertex, 0

G
vvertex . Thus, the equation set suitable for KIVA
It should be noted that Eq. (3) is for the fully developed
implementation is:
turbulent flame, and the relation l F , t = b 2 l was used in
i G G the derivation by Peters [2], where b2 = 1.78 is a constant
∂G i = ρu S 0 | ∇G
+ (vjf − vvertex ) ⋅∇G i | − D κi | ∇G
i| (1)
∂t ρ
T t obtained from dimensional reasoning. It was shown by
Peters [2] and Ewald [9] that the unsteady solution of lF ,t
k can be derived from Eq. (2) by assuming that the
∂G ′′2 G G k ρ k
+ (v f − vvertex ) ⋅ ∇G ′′2 = ∇ & ⋅ ( u Dt ∇ & G ′′2 ) turbulence quantities Dt , k , and ε are constant, and by
∂t ρ
(2) assuming a uniform turbulence profile. Based on these
 2
+ 2 Dt (∇G i )2 − c ε G k ′′ assumptions, the convection and diffusion terms which
s 
k k
include the gradient of G ′′2 all vanish, and Eq. (2) can be
reduced to an ordinary differential equation for the
turbulent flame brush thickness (cf. [2]):
dl F , t Although the above G-equation description was
= b22 c s l 2 − c s l F2 , t (8) originally developed for premixed flames by Peters [2], it
d (t / τ )
is also applied to partially premixed flames in Direct
Injection SI (DISI) engines in this study. This scenario
where τ = k ε is used as a
nondimensionalizing features the so-called triple flame structure, as shown in
Fig. 1. The premixed flame branches are described
timescale, and correlations relating Dt , k , and ε to u ′ ,
using the G-equation, while the secondary heat release
l , and τ have been used. Choosing lF ,t = 0 as the initial and pollutant formation within the diffusion flames
value at spark timing, an algebraic solution results: behind the flame front are modeled by detailed chemical
kinetics.
l F ,t = b2 l[1 − exp( − cs t / τ )]1 2 (9)
The prediction of the turbulent burning velocity plays a
crucial role in the modeling of SI engine combustion.
Based on Eq. (9), the unsteady turbulent flame speed is, The laminar flame speed is one of the most important
scaling factors in most of the published correlations for
⎧ ⎤ ⎫⎪
12
⎪ a4b3 l ⎡⎛ a4b3 l ⎞ the turbulent flame speed. Metgalchi et al. [17] found
2
2 u ′l
2 2
ST0
= 1 + I P ⋅ ⎨− + ⎢⎜ ⎟ + a4b3 0 ⎥ ⎬ (10) that the experimentally measured laminar flame speed
SL0 ⎪⎩ 2b1 lF ⎢⎣⎝ 2b1 lF ⎠ sL lF ⎥ ⎪
⎦ ⎭ can be correlated as a function of equivalence ratio,
temperature and pressure by:
where
α β
I P = [1 − exp(−cs t / τ )]1 2 (11) ⎛ Tu ⎞ ⎛ P ⎞
S =S
0
L
0
L , ref ⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟ ⋅ Fdil (13)
⎝ Tu , ref ⎠ ⎝ Pref ⎠
Physically, the additional exponential term, I P , can be
interpreted as a progress variable which accounts for the where the subscript ref means the reference condition of
increasing disturbing effect of the surrounding eddies on 298K and 1atm. Fdil is a factor accounting for the
the flame front surface as the ignition kernel grows from
diluent’s effect. The fuel-type independent exponents α
the laminar flame stage into the fully developed turbulent
and β were correlated as functions of equivalence ratio
stage. Similar terms also appear in the turbulent flame
speed correlation for an ignition kernel flame by Herweg as:
et al. [16]. According to Peters [2], cs = 2.0 is a constant
derived from spectral closure. However, in practical α = 2.18 − 0.8(φ − 1) (14)
engine simulations, uncertainties associated with other
sub-models including chemistry mechanisms or even β = −0.16 + 0.22(φ − 1) (15)
mesh resolution during the kernel growth could result in
difficulties in matching experimental data. Therefore, this The reference flame speed is given as:
progress variable was selected to be tunable in the
present study by introducing a model constant, Cm 2 , S L0 , ref = B M + B2 (φ − φ M ) 2 (16)
while keeping the same scaling relation, i.e.,
Values for BM , B2 and φM for propane and isooctane are
I P = [1 − exp(−t /(Cm 2 ⋅τ ))]1 2 (12)
listed in Table 1.

Although Cm 2 is regarded as tunable for different Table 1 Values for BM , B2 and φM with Eq. (16)
engines, it is fixed for all spark timing sweeps, rpm Fuel BM (cm/s) B2 (cm/s) φM
sweeps and Exhaust Gas Recirculation (EGR) sweeps
for each specific engine in the present work. Propane 34.22 -138.65 1.08
Isooctane 26.32 -84.72 1.13

Unfortunately, Eq. (16) predicts negative flame speeds


for very lean or very rich mixtures. For example, the
flammability range of isooctane is generally quoted as
0.6 < φ < 1.7 , which is acceptable for simulations of
premixed flames near stoichiometric conditions, but is
not applicable to stratified charge combustion in DISI
engines. As suggested by Deur et al. [18], one practical
solution is to follow the expression proposed by Gülder
Figure 1 Triple flame structure in DISI engines [11]. [19] in which the flame speed will never be driven
negative, viz.,
S L0 , ref = ωφ η exp −ξ (φ − σ( )
2
) (17) Fdil = 1 − 2.06 ⋅ X dil
0.773
(19)

where X dil is the mole fraction of the diluent.


where ω , η , ξ , and σ are data fitting coefficients.
However, the reference flame speeds based on the
In this study, we combined the correlations by Ryan [21]
values of the coefficients originally suggested by Gülder
and by Metghalchi et al. [20], i.e., in Eq (18)
[19] show a relatively large discrepancy with published
experimental data in the literature [17]. For isooctane,
following the idea of Deur [18], in the present study, a f = 2.1 + 1.33 ⋅ Ydil for 0 < Ydil < 0.2 (20)
group of new values of the coefficients in Eq. (17) was
obtained by correlating the data of Metgalchi et al. [17] f = 2.5 for 0.2 < Ydil < 0.476 (21)
within the range 0.65 < φ < 1.6 . Gülder’s values and the
new values are listed in Table 2, and a comparison of However, it can be seen that this expression fails
isooctane reference flame speeds from the different beyond Ydil > 0.476 because the flame speed becomes
correlations is shown in Fig. 2. The same treatment can negative outside this range. However, the new
be adopted for other fuel types. correlation gives good results in engine simulations as
will be shown later. The different mass fraction-based
Table 2 Coefficient values in Eq. (17) for isooctane diluent factors are compared in Fig. 3. In comparison,
the mole fraction-based Eq. (19) results in more
ω η ξ σ reduction effect compared to all the mass fraction-based
Gülder [19] 46.58 -0.326 4.48 1.075 correlations presented here.
Present 26.9 2.2 3.4 0.84
50 1.0 Metgalchi et al.
Metghalchi et al. Ryan et al.
Gulder
Present study
40 Present study
0.8
SL,ref [ cm/sec ]

30 0.6
Fdil
0

20 0.4

10 0.2

0 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35


0.0 0.5 1.0 1.5 2.0 2.5 Ydil
φ
Figure 3 Comparison of different diluent factors.
Figure 2 Comparison of laminar flame speed
correlations for isooctane under reference conditions
Primary heat release within the turbulent flame brush
(T=298K, P=1atm).
In the present implementation of the G-equation model,
The presence of diluent due to internal residual and/or
it is assumed that after the flame front has passed, the
EGR has a significant effect on the laminar flame speed.
mixture within the mean flame brush tends to the local
This effect is usually accounted for by a term such as
and instantaneous thermodynamic equilibrium. The
Fdil in Eq. (13). One expression suggested for all fuel
species conversion rate and the associated primary heat
types is [20]: release at the flame front are calculated based on this
assumption.
Fdil = 1 − f ⋅ Ydil (18)
Tan and Reitz [12] suggested a method for calculating
where Ydil is the mass fraction of diluent and f is an the species density change in the cells containing the
experimentally determined constant. Ryan et al. [21] mean flame front, viz.,
suggested that f = 2.5 is valid for 0 < Ydil < 0.3 .
d ρi Af ,i 4 0
Metgalchi et al. [20] suggested f = 2.1 to be valid for = ρ (Yi ,u − Yi ,b ) ST (22)
dt Vi 4
0 < Ydil < 0.2 . Rhodes et al. [22] proposed another
expression based on laminar flame speed
where ρ is the average density of the mixture in cell,
measurements of indolene-air-diluent mixtures, viz.,
i 4 . Yi ,u and Yi ,b are the mass fractions (i.e., fractions of
the total mass in the cell) of species i in the unburnt and The steps to determine the unburnt species mass
burnt mixtures, respectively. Af is the mean flame front fractions Yi ,u are as follows:
area and V is the cell volume. i 4 is the cell index used
in KIVA. Tan and Reitz [12] assumed seven species in 1. Determine the equilibrium species mass fractions
their study, including fuel, O2, N2, CO2, H2O, CO, H2. Yi ,b and the adiabatic flame temperature Tb = Tadia . In
Yi ,b was also assumed to be the local equilibrium value. this study, a Fortran code by Pope [23,24] was used
One issue associated with Eq. (22) is how to determine for the calculation of the chemical equilibrium. This
Yi ,u . By assuming the Yi ,u of N2, CO2, H2O, CO, H2 to be code is based on the element potential method as in
the STANJAN code [25], but an improved numerical
constantly equal to their initial values at the time of
algorithm called Gibbs function continuation is
ignition, the Yi ,u of fuel and O2 can be estimated from the
adopted for better computational stability and
C, H and O element conservation relations (cf. Tan [8]). efficiency.
However, when a large number of intermediate species 2. Calculate the burnt gas density and the burnt
are included, and detailed chemistry is considered as in species densities based on the equation of state and
the present study, the above method is no longer able to the mass fractions Yi ,b from step 1.
give the predicted Yi ,u for all the species. This is
because the number of elements and therefore the Pi 4 ⋅ MWmix ,b
available conservation equations is limited. Therefore, a ρb = (25)
new method is suggested that is based on the sub-grid Ru Tb
scale unburnt/burnt volumes of the flame-containing
cells. In this method, it is assumed that the mean flame ρi ,b = ρb ⋅ Yi ,b (26)
front surface cuts every flame-containing cell into two
parts, an unburnt volume Vu and a burnt volume Vb , as where MWmix ,b is the average molecular mass of the
shown in Fig. 4. As the mean flame front sweeps
burnt mixture, and Ru is the universal gas constant.
forward, the mixture within the sweeping volume tends
to local equilibrium following a constant pressure,
constant enthalpy process. The pressure is assumed to 3. Calculate the unburnt species densities ρi ,u based
be homogeneous across the flame in the cell, consistent on species mass conservation.
with deflagration wave theory. The sub-grid scale
volumes are tracked for every time step based on the ρiVi 4 − ρi ,bVb
coordinate information of the cell vertices and the flame ρi ,u = (27)
surface piercing points. The species density conversion Vu
rate then becomes:
4. Finally, determine the unburnt species mass
d ρi Af ,i 4 0 fractions:
= ρu (Yi ,u − Yi ,b ) ST (23)
dt Vi 4
ρi ,u
Yi ,u = (28)
Compared to Eq. (22), the cell averaged density ρ is ∑ ρi ,u
i
replaced by the density of the unburnt mixture ρu and
now the Yi ,u and Yi ,b are evaluated with respect to the In KIVA, the heat release due to the chemistry source
mass of unburnt and burnt mixture, respectively, viz., term is directly related to the species conversion [26]. It
needs to be noted that the four species associated with
the NOX formation mechanism, i.e., NO, NO2, N, and
∑Y
i
i ,u =∑ Yi ,b =1
i
(24) N2O [27], are excluded from the equilibrium calculation
due to their relatively short residence time within the
Unburnt flame front, and the relative slow rate of the NOX
chemical reactions.
Vu
Post-flame heat release and pollutant formation
Vb
Burnt Fundamental understanding of the chemical processes
occurring behind the turbulent flame brush in SI engines
Mean Flame Front
is still incomplete, especially for partially premixed
combustion. Whether the species conversion in this
region is turbulence mixing-controlled or chemical
Figure 4 Numerical descriptions of the turbulent flame kinetics-controlled should be answered by further
structure and the flame containing cells. experimental or Direct Numerical Simulation (DNS)
investigations. In this study, the computational cells
behind the flame front are modeled as Well Stirred IGNITION KERNEL MODEL
Reactors (WSR). Detailed hydrocarbon oxidation
chemical kinetic mechanisms are applied to account for The growth of the ignition kernel is tracked by using the
the further oxidation of CO and other intermediate DPIK model by Fan, Tan, and Reitz [11,31]. By
species, such as small hydrocarbon molecules and the assuming a spherical shaped kernel, the flame front
species in the H2-O2 system. To consider the effects of position is marked by Lagrangian particles, and the
turbulent mixing, the reaction rates can be adjusted by flame surface density is obtained from the number
considering the eddy turnover time as a turbulent density of particles in each computational cell. Assuming
timescale, and by combining this timescale with a kinetic the temperature inside the kernel to be uniform, the
timescale. This kinetic timescale can be selected as the kernel growth rate is:
conversion timescale of a heat-release-rate-limiting
species, such as CO (cf. Kong et al. [28]). drk ρ
= u ( S plasma + S T ) (30)
dt ρk
A nine-reaction reduced NOX mechanism was coupled
with the hydrocarbon oxidation mechanism for predicting where rk is the kernel radius, ρ u is the local unburnt
the formation of NO and NO2. The reactions include [27]:
gas density, and ρ k is the gas density inside the kernel
N + NO ⇔ N 2 + O (29a) region.
N + O2 ⇔ NO + O (29b)
The plasma velocity S plasma is given as [8]:
N 2 O + O ⇔ 2 NO (29c)
N 2 O + OH ⇔ N 2 + HO2 (29d) Q spk ⋅ η eff
N 2 O(+ M ) ⇔ N 2 + O(+ M ) (29e) S plasma = (31)
⎡ ρ ⎤
HO2 + NO ⇔ NO2 + OH (29f) 4π rk2 ⎢ ρ u ( u k − hu ) + P u ⎥
⎣ ρ k ⎦
NO + O + M ⇔ NO2 + M (29g)
NO2 + O ⇔ NO + O2 (29h) where Q spk is the electrical energy discharge rate, η eff
NO2 + H ⇔ NO + OH (29i) is the electrical energy transfer efficiency due to heat
loss to the spark plug. η eff = 0.3 , as suggested by
Chemical kinetics based end-gas autoignition modeling Heywood [32] is used in this study. ρ u and hu are the
density and enthalpy of the unburnt mixture. ρ k and u k
It is generally accepted that engine knock is caused by
the autoignition of a portion of the end-gas prior to the are the density and internal energy of the mixture inside
flame arrival. A simple but widely used approach to the kernel.
predict engine knock is the Livengood-Wu integral [29],
which essentially uses a one-step reaction to The laminar flame speed S L0 in Eq. (10) was multiplied
approximate the autoignition mechanism. Recently, by a stretch factor, I 0 , which accounts for strain and
researchers have proposed using more detailed
curvature effects, and the modified correlation is used as
mechanisms to predict the autoignition time more
the turbulent flame speed, S T , in the kernel stage. I 0
accurately. For example, Eckert et al. [30] applied the
“Shell” autoignition model to simulate the autoignition of follows the suggested expression of Herweg et al. [16]:
the end-gas in SI engines. Research on detailed
3 2
chemical kinetic mechanisms for the autoignition of ⎛ l ⎞
1 2
⎛ u′ ⎞ lF ρ u
fuel/air mixtures has achieved much more accurate I0 = 1 − ⎜ F ⎟ ⎜ 0⎟ − 2⋅ (32)
⎝ 15l ⎠ ⎝ SL ⎠ rk ρ k
predictions of the autoignition delay time and the related
thermo-chemical parameters. This makes it possible to Note that curvature effects are also considered in the
accurately describe the location and intensity of the end- present combustion model by the last term of Eq. (1).
gas autoignition in SI engines without tuning any
reaction rate constants. In this study, detailed chemical The chemistry processes in the kernel growth stage are
kinetic mechanisms are also applied in each cell in front treated in the same way as in the G-equation
of the mean flame front. The auto-ignited mixture can be combustion model. Although the transport equation of
tracked by monitoring certain species that characterize i is not solved here, the G
i field is constructed based
the high temperature heat release, such as the OH G
radical. Although no knocking modes are considered in on the positions of the kernel particles, thus providing
the test cases in this paper, this methodology will be the necessary information for the chemical heat release
followed in our future study of engine knock. calculations.

The transition from the kernel model to the turbulent G-


equation model follows the same criterion as the one
used in the previous work by Tan and Reitz [36],
namely, that the transition is controlled by a comparison Table 4 Caterpillar C3H8 engine specifications.
of the kernel radius with a critical size which is
proportional to the locally averaged turbulence integral Engine CAT 3401 SI Gas
length scale, viz., Retrofit
Bore × Stroke (mm) 137.16 × 165.1
k3 2
rk ≥ C m 1 ⋅ l = C m 1 ⋅ 0.16 (33) Compression Ratio 10:1
ε
Equivalence Ratio Stoichiometric
where C m 1 is a model constant. Compared with the
previous work by Tan [8], where two different turbulent Spark Duration (ms) 2
flame speed correlations were applied in the kernel Intake Valve Closure (˚) -147 ATDC
model and in the G-equation combustion model, C m 1 is Initial Pressure at IVC (kPa) 51.0 kPa
no longer as crucial in the model calibration since the
turbulent flame speed correlations used in the kernel
model and the G-equation combustion model are Table 5 Operational settings and measured initial
essentially consistent. conditions for the Caterpillar C3H8 engine [34].
Engine Speed/ Ignition Residual Temperature
ENGINE SIMULATION RESULTS % EGR Timing Fraction at IVC
(˚ATDC) (%) (K)
Two SI engines were modeled as test cases for the 1600 (rev/min) -10
present combustion model. One is a homogeneous 0% EGR -20 14.5 367.8
charge Caterpillar converted propane-fueled engine; the -30
other is a two-stroke Mercury Marine DISI gasoline -40
engine. In the simulations, CHEMKIN II [33] was used 1600 (rev/min) -10 14.5 361.7
for solving the detailed chemical kinetics. 5% EGR
1600 (rev/min) -10 14.5 362.0
CATPILLAR-CONVERTED PROPANE ENGINE 10% EGR
1000 (rev/min) -10 17 339.4
The experimental data for the homogeneous charge 0% EGR
Caterpillar propane engine cover different operating 1900 (rev/min) -10 13 346.9
conditions, including a spark timing sweep, an EGR level 0% EGR
sweep and an engine speed sweep [34]. The engine
specifications and operating conditions are listed in The model constants C m 1 = 2.0 and C m 2 = 1.0 were
Table 4 and Table 5. A 100-species, 539-reaction
propane mechanism from the Lawrence Livermore used in all simulated cases for this engine. As
National Lab (LLNL) was used in the calculations [10]. mentioned earlier, C m 2 is the only crucial constant to be
calibrated.
Figure 5 shows a 360˚ mesh and a 2D sector mesh of
the axially symmetric cylinder. The 2D sector mesh with Figure 6 shows comparisons of measured and predicted
periodic boundaries was used due to CPU time in-cylinder pressure and engine-out NOX for different
considerations. The spark plug is located at the center of spark timings (two groups of repeatedly measured NOX
the cylinder head, and is represented using stationary data are shown in 6b). As seen, the predicted results
particles, as implemented by Fan [31]. The initial in- match the measured data reasonably well.
cylinder mixture was assumed to be completely
homogeneous. The simulations start from IVC. The The effect of EGR level on the combustion is shown in
initial swirl ratio was set as 0.78 according to Fig. 7. An increase of EGR ratio slows down the
experimental measurements. The initial turbulence combustion by reducing both the laminar and turbulent
kinetic energy was assumed to be uniform in the flame speeds (cf. Eq. (13)). The NOX emissions also
cylinder, with an estimated value equal to the kinetic decrease with increased EGR due to the reduction of
energy based on the mean piston speed. peak temperature. The model captures the general
trends very well, as shown by the good matching
between the measured and predicted data. It needs to
be noted that the total residual fraction in Eq. (18)
includes both the EGR and the estimated internal
residual (14.5%).

Figure 8 shows the pressure traces and NOX data for


different engine speeds. The predicted pressure curves
360˚ full mesh 2D sector mesh generally match the experimental pressures reasonably
Figure 5 Computational meshes of Caterpillar converted well. The relatively large discrepancy between the
propane SI engine at -30˚CA ATDC. measured and predicted NOX in the 1000rpm case may
3.5 4000
O
3.0 Measured - 40 ATDC 3500
Predicted
2.5 - 30
O
ATDC 3000
Pressure [ MPa ]

2500

NOX ( ppm )
2.0 - 20
O
ATDC
2000
1.5 O
- 10 ATDC 1500
1.0 Measured Data1
1000
Measured Data2
0.5 500 Predicted

0.0 0
-100 -80 -60 -40 -20 0 20 40 60 80 100 -50 -40 -30 -20 -10 0
O
O
Spark Timing ( CA ATDC )
Crank Angle [ ATDC ]
(6a) (6b)
Figure 6 In-cylinder pressure curves (6a) and engine-out NOX (6b) for the spark timing sweep. (EGR=0%, engine
speed=1600 rev/min, residual fraction=14.5%).

3.5
4000
Measured
3.0 0 % EGR
Predicted 3500

2.5 5 % EGR 3000


Pressure [ MPa ]

2.0 10 % EGR 2500


NOX [ ppm ]

2000
1.5 Measured
1500 Predicted
1.0
1000
0.5
500
0.0 0
-100 -80 -60 -40 -20 0 20 40 60 80 100 0 2 4 6 8 10
O
Crank Angle [ ATDC ] EGR [ % ]
(7a) (7b)
Figure 7 In-cylinder pressure curves (7a) and engine-out NOX (7b) for the EGR level sweep. (Engine speed=1600
rev/min, spark timing=-40˚ ATDC, residual fraction=14.5%).

3.0 2400

2.5 Measured
2000
Predicted
1000 rpm
2.0 1600
Pressure [ MPa ]

NOX [ ppm ]

1.5
1200
1600 rpm
1.0
1900 rpm 800
Measured
0.5 Predicted
400

0.0
0
-100 -80 -60 -40 -20 0 20 40 60 80 100 1000 1200 1400 1600 1800 2000
O
Crank Angle [ ATDC] Engine Speed [ RPM ]
(8a) (8b)
Figure 8 In-cylinder pressure curves (8a) and engine-out NOX (8b) for the engine speed sweep.(Spark timing=-10˚ ATDC,
EGR=0%).
be due to the under-predicted peak pressure (and
therefore peak temperature) compared to the
experimental pressure.

Figure 9 shows the simulated in-cylinder temperature


profiles as the flame propagates out from the spark plug,
where the black contour line denotes the locations of the
mean flame front ( G i = 0 iso-surface). As can be seen,
the temperature of the mixture immediately behind the
turbulent flame brush is above 2500K, which is
approximately equal to the local equilibrium temperature.

Figure 10 shows the predicted in-cylinder species mass


fractions of C3H8, CO2, CO, OH, NO and NO2 at -6˚ CA
ATDC for the spark timing=-30˚ ATDC case. After the
flame front passes, the parent fuel molecules are
consumed. The subsequent chemistry process behind
the flame brush is governed by the CO oxidation
reactions, the H2-O2 system reactions and the NOX
formation mechanism. The mass fraction of NO reaches
its peak value in the highest temperature region, as
Figure 10 Species mass fractions at CA=-6˚ ATDC
expected. Most of the NO2 is generated right ahead of
(Spark timing=-30˚ ATDC).
the mean flame front. This is because NO2 formation is
favored under relatively low temperature combustion
conditions. Considering that the peak mass fraction of
NO2 is two orders of magnitude less than that of NO, the
MERCURY MARINE TWO-STROKE GDI ENGINE
NO2 emission is essentially negligible in this case.
To further test the applicability of the models to partially
premixed flames, a two-stroke marine gasoline DISI
engine was also modeled. The specifications and tested
conditions are listed in Table 6 [35]. The computing
mesh at -77˚ ATDC is shown in Fig. 11. The direct
injection combustion system of this engine employs an
air-guided scheme. The engine has a flat-top piston and
a high-pressure swirl-type injector that is centrally
mounted in the dome-shaped cylinder head. The in-
cylinder tumble flow transports the injected fuel to the
spark plug that is located on the exhaust port side. A
stratified mixture is formed around the spark gap at the
time of the spark.

The simulations start from exhaust port open (-262˚


ATDC) and cover a full two-stroke engine cycle. The
initial swirl ratio was set as 0 because all the ports were
still closed at this crank angle. The initial cell turbulence
kinetic energy density in the cylinder and all the ports
was assumed to be uniform and estimated to be equal to
the kinetic energy density based on the mean piston
speed. An inflow boundary condition was specified at the
inlet of the boost port, where the experimentally
measured crank case pressures were specified versus
crank angles. The outlet of the exhaust port was set as
an open boundary, where the measured exhaust
pressures were specified versus crank angles. The
scavenging and mixture formation processes have been
studied by Tan and Reitz [36] in previous work, and we
Figure 9 In-cylinder temperature contours at different focus only on the combustion phase in this study.
crank angles (Spark timing=-30˚ ATDC). Solid black line
indicates the flame front location.
3.0
Measured
Location of Predicted
O
-41 ATDC
2.5
O
-31 ATDC
Spark Plug
2.0

Pressure (MPa)
O
-21 ATDC
1.5
Exhaust Port -16 ATDC
O

1.0
Transfer Port
0.5

Boost Port
0.0
-100 -50 0 50 100
O
Crank Angle ( ATDC)
Figure 11 Computational mesh of the Mercury Marine
two-stroke gasoline direct injection engine. (12a)

Table 6 Mercury Marine two-stroke GDI engine 800


specifications and experimental conditions [35]. 700
Measured
Predicted
Bore × Stroke (mm) 85.8 × 67.3 600

Compression Ratio 7.4:1 500

Exhaust Port Timing (˚ ATDC) 95 NOX (ppm) 400


Boost Port Timing (˚ ATDC) -243 300
Transfer Port Timing (˚ ATDC) -243
200
Injection Timing (˚ ATDC) -85
100
Engine Speed (rev/min) 2000
0
Spark Timing (˚ ATDC) -41, -31, -21, -16 -45 -40 -35 -30 -25 -20 -15
Total Sprayed Fuel Mass (g) 7.29e-3 O
Spark Timing ( CA ATDC)
(12b)

To simulate the chemical kinetics of gasoline Figure 12 In-cylinder pressure curves (12a) and engine-
combustion, a 21-species, 42-reaction isooctane out NOX (12b) of Mercury Marine engine for spark timing
(iC8H18) mechanism was used. This mechanism is sweep (Engine speed=2000 rev/min).
extracted from the reduced Primary Reference Fuel
(PRF) mechanism of Tanaka et al. [37] by excluding the The development of the simulated flame front surface
n-heptane (nC7H16) related species and reactions. Four based on the updated laminar flame speed correlation
cases were simulated with a spark timing sweep. The (Eq. (17)) was also compared with the results using the
simulated pressure traces and engine-out NOX were previous correlation, Eq. (16), as shown in Fig. 13. It can
compared with the measured data. In the simulations of be seen that the left side of the flame front becomes
this engine, the model constants C m 1 = 2.0 and nearly stagnant after around 33˚ ATDC in the pictures in
the left column. This is because the equivalence ratios of
C m 2 = 3.0 were fixed and no case-by-case adjustment
the mixture to the left side of this flame front branch are
was made. beyond the flammability limits predicted by Eq. (16).
Physically, this prediction is not reasonable because it is
Figure 12a shows the comparison of in-cylinder well known that isooctane/air mixtures leaner than
pressures. The matching between the measured [35] φ = 0.65 are able to be ignited. In contrast, the predicted
and predicted peak pressures and combustion phasing results of Eq. (17) look more realistic. Although more
is reasonably good considering the complicated spray fundamental study needs to be done for more accurate
and flow conditions in this two-stroke configuration. The correlations of the flame speed of lean/rich mixtures, the
predicted NOX data are compared with the measured improvement of Eq. (17) over Eq. (16) is considered to
data in Fig. 12b. Although there are discrepancies in the be necessary and considerable.
absolute values, the general trends match pretty well.
CONCLUSION

Detailed chemical kinetics was coupled with a G-


equation combustion model for better predictions of fuel
oxidation, pollutant formation. The integrated model was
implemented into the KIVA-3V CFD code to simulate
both homogeneous charge and stratified charge
combustion in SI engines. A Caterpillar-converted
homogeneous charge propane SI engine and a Mercury
Marine GDI engine were modeled. The predicted in-
cylinder pressure traces and engine-out NOX were
compared with the measured data and good agreement
was achieved for a wide range of operating conditions.
The in-cylinder temperature and species mass fraction
contours of selected cases were also shown and
explained.

Previously proposed turbulent flame speed correlations


were restudied and updated in the present model for a
Figure 13 Temperature contour evolution in the Mercury better description of the ignition kernel flame evolution.
Marine engine showing the influence of the laminar The laminar flame speed correlations were modified to
flame speed correlations (Left column: predicted using account for very lean and very rich mixtures under
Eq. (16); Right column: predicted using Eq. (17)). partially premixed combustion conditions. The diluent
effect on the flame speed was also reconsidered for
better matching with the experimental data.

A new method based on the sub-grid scale


unburnt/burnt volumes was suggested for the species
conversion and primary heat release calculations within
the turbulent flame brush. In this method, a numerically
improved code based on the element potential method
was used for the equilibrium calculations. This method
proved to be effective of dealing with the consideration
of large number of species.

Methodologies of using detailed chemistry for the post-


flame heat release and pollutant formation calculations,
and for end-gas autoignition predictions were also
introduced. More emphasis will be placed on them in
future work.

ACKNOWLEDGMENTS

Ford Motor Company is greatly acknowledged for the


financial support of this project. The authors also thank
Dr. Youngchul Ra at the Engine Research Center for
Figure 14 Species mass fractions and flow velocities at helpful discussions.
35˚ ATDC (spark timing=-31˚ ATDC).
REFERENCES
Figure 14 shows the species mass fractions of iC8H18,
O2, CO, OH, NO and velocity contours on a clip plane
1. Abraham, J., Bracco, F. V., and Reitz, R. D.,
through the cylinder diameter at 35˚ ATDC (spark
“Comparisons of Computed and Measured
timing=-31˚ ATDC). As in Fig. 9, the black lines
represent the mean flame font location. At this time, the Premixed Charge Engine,” Combustion and Flame,
fuel within the burnt region has all broken down to 60: 309-322, 1985.
smaller species. Most of the CO is generated behind the 2. Peters, N., Turbulent Combustion, Cambridge
left branch of the flame font because some fuel and O2 University Press, 2000.
are convected across the flame front from the left side 3. Bray, K. N. C., Libby, P. A., “Recent Developments
by the in-cylinder flow, which can be seen from the in the BML Model of Premixed Turbulent
velocity contours. Most of the OH and NO occur within Combustion,” Turbulent Reacting Flow, Academic
the high temperature burnt region. Press, London, 1994.
4. Boudier, P., Henriot S., Poinsot, T., and Baritaud, T., 19. Gülder, O. L., “Correlations of Laminar Combustion
“A Model for Turbulent Flame Ignition and Data for Alternative S.I. Engine Fuels,” SAE Paper
Propagation in Spark Ignition Engines,” 24th 841000, 1984.
Symposium (International) on Combustion/ The 20. Metghalchi, M., and Keck, J. C., “Burning Velocities
Combustion Institute, pp. 503-510, 1992. of Methanol, Ethanol and Iso-octane-Air Mixtures”,
5. Williams, F. A., Turbulent Combustion, SIAM, 19th Symposium (International) on Combustion/ The
Philadelphia, 1985. combustion Institute, pp. 275, 1983.
6. Peters, N., “The Turbulent Burning Velocity for Large 21. Ryan, T. W. and Lestz, S. S., “The Laminar Burning
Scale and Small Scale Turbulence,” J. Fluid Mech., Velocity of Iso-octane, n-heptane, Methanol,
384: 107-132, 1999. Methane, and Propane at Elevated Temperature
7. Dekena M., and Peters N., “Combustion Modeling and Pressure in the Presence of a Diluent,” SAE
with the G-equation,” Oil & Gas Science and Paper 800103, 1980.
Technology-Rev. IFP, 54(2): 265-270, 1999. 22. Rhodes, D. B. and Keck, J., “Laminar Burning
8. Tan. Z., Multi-Dimensional Modeling of Ignition and Speed Measurements of Indolene-Air-Diluent
Combustion in Premixed and DIS/CI (Direct Injection Mixtures at High Pressures and Temperatures,” SAE
Spark/Compression Ignition) Engines, Ph.D. Thesis, Paper 850047, 1985.
University of Wisconsin-Madison, 2003. 23. Pope, S. B., “The Computation of Constrained and
9. Ewald, J. and Peters, N., “A Level Set Based Unconstrained Equilibrium Compositions of Ideal
Flamelet Model for the Prediction of Combustion in Gas Mixtures Using Gibbs Function Continuation”,
Spark Ignition Engines,” 15th International FDA03-02, Cornell University, http://eccentric.mae.
Multidimensional Engine Modeling User’s Group cornell.edu/~pope/Reports/CEQ_FDA.pdf, 2003.
Meeting, Detroit, MI, 2005. 24. Pope, S. B., “CEQ: A Fortran Library to Compute
10. Curran, H. J., Gaffuri, P., Pitz, W. J., and Westbrook, Equilibrium Compositions Using Gibbs Function
C. K., “A Comprehensive Modeling Study of Iso- Continuation,”
octane Oxidation,” Combustion and Flame, 129(3): http://eccentric.mae.cornell.edu/~pope
253-280, 2002. /CEQ, 2003.
11. Tan, Z., Kong S. C., and Reitz R. D., “Modeling 25. Reynolds, W. C., “The Element Potential Method for
Premixed and Direct Injection SI Engine Combustion Chemical Equilibrium Analysis: Implementation in
Using the G-Equation Model,” JSAE/SAE the Interactive Program STANJAN,” Stanford
International Spring Fuels & Lubricants Meeting, University, Dept. of Mechanical Engr., 1986.
Japan, JSAE Paper 2003-01-1843, 2003. 26. Amsden, A. A., “A Block-structured KIVA Program
12. Tan, Z. and Reitz R. D., “Modeling Ignition and for Engines with Vertical or Canted Valves,” Los
Combustion in Spark-ignition Engines Using a Level Alamos National Lab Report, LA-13313-MS, 1997.
Set Method,” SAE Paper 2003-01-0722, 2003. 27. Kong, S. C., Sun, Y., and Reitz, R. D., “Modeling
13. Han, Z. and Reitz, R. D., “Turbulent Modeling of Diesel Spray Flame Lift-off, Sooting Tendency and
Internal Combustion Engines Using RNG k-e NOX Emissions Using Detailed Chemistry with
Models,” Comb. Sci. Tech. 106: 267-295, 1995. Phenomenological Soot Model,” Proceedings of
14. Göttgen, J., Mauss, F., and Peters, N., “Analytic ICES2005, ASME ICED 2005 Spring Technical
Approximations of Burning Velocities and Flame Conference, Chicago, IL, 2005.
Thickness of Lean Hydrogen, Methane, Ethylene, 28. Kong, S. C., Marriott, C. D., Reitz, R. D., and
Ethane, Acetylene, and Propane Flames,” 24th Christensen, M., “Modeling and Experiments of
Symposium (International) on Combustion/ The HCCI Engine Combustion Using Detailed Chemical
combustion Institute, pp. 129-135, 1992. Kinetics with Multidimensional CFD,” SAE 2001-01-
15. Smooke, M. D. and Giovangigli, V., Reduced 1026, 2001.
Kinetic Mechanisms and Asymptotic Approximations 29. Livengood, J. C. and Wu, P. C., “Correlation of
for Methane-Air Flames, Lecture Notes in Physics, Autoignition Phenomenon in Internal Combustion
384, pp. 1, Springer, Berlin, 1991. Engines and Rapid Compression Machines,” 5th
16. Herweg, R. and Maly, R. R., “A Fundamental Model Symposium (International) on Combustion/ The
for Flame Kernel Formation in S.I. Engines,” SAE Combustion Institute, pp. 347-356, 1955.
922243, 1992. 30. Eckert, P., Kong, S. C., and Reitz, R. D., “Modeling
17. Metghalchi, M., and Keck, J. C., “Burning Velocities Autoigniton and Engine Knock Under Spark Ignition
of Mixtures of Air with Methanol, Isooctane, and Conditions,” SAE Paper 2003-01-0011, 2003.
Indolene at High Pressures and Temperatures,” 31. Fan, L., and Reitz, R. D., “Development of Ignition
Combustion and Flame, 48: 191-210, 1982. and Combustion Model for Spark-Ignition Engines,”
18. Deur, J. M., Raghunathan, B., and Dhanapalan, S., SAE Paper 2000-01-2809, 2000.
“Combustion and Spray Simulation of a GDI Engine 32. Heywood, J. B., Internal Combustion Engine
Incorporating a Multi-Component Fuel,” 10th Fundamentals, McGraw-Hill, 1988.
International Multidimensional Engine Modeling 33. Kee, R. J., Rupley, F. M., and Miller, J. A.,
User’s Group Meeting, Detroit, MI, 2000. “CHEMKIN-II: A FORTRAN Chemical Kinetics
Package for the Analyses of Gas Phase Chemical CONTACT
Kinetics,” Sandia Report, SAND 89-8009, 1989.
34. Hampson, G. J., A Theoretical and Experimental Long Liang
Study of Emission Modeling for Diesel Engines with Engine Research Center
Comparison to In-cylinder Imaging, Ph.D. Thesis,
University of Wisconsin-Madison
University of Wisconsin-Madison, 1997.
35. Hudak, E., Time-Resolved Exhaust Measurements 1500 Engineering Drive, Madison WI 53706, USA
of a Two-Stoke Direct-Injection Engine, Master lliang@wisc.edu
Thesis, University of Wisconsin-Madison, 1998.
36. Tan, Z., Fan, L. and Reitz, R. D., “Modeling Ignition,
Multi-component Fuel Vaporization and Spray ABBREVIATIONS
Breakup in a DISI Engine,” ILASS Americas, 14th
Annual Conference on Liquid Atomization and Spray ATDC After Top Dead Center
System, Dearborn, MI, May, 2001. Characteristic Timescale Combustion (Model)
CTC
37. Tanaka, S., Ayala, F., and Keck, J. C., “A Reduced
Chemical Kinetic Model for HCCI Combustion of DISI Direct Injection Spark Ignition
Primary Reference Fuels in a Rapid Compression DPIK Discrete Particle Ignition Kernel
Machine,” Combustion and Flame, 133: 467-481, GDI Gasoline Direct Injection
2003. EGR Exhaust Gas Recirculation
IVC Intake Valve Closure
PDF Probability Density Function
PRF Primary Reference Fuel
SI Spark Ignition

View publication stats

Anda mungkin juga menyukai