You are on page 1of 16

SPECIAL ISSUE PAPER

497

A fully coupled computational uid dynamics and


multi-zone model with detailed chemical kinetics for
the simulation of premixed charge compression ignition
engines
A Babajimopoulos1*, D N Assanis1, D L Flowers2, S M Aceves2, and R P Hessel3
1Department of Mechanical Engineering, University of Michigan, Ann Arbor, Michigan, USA
2Lawrence Livermore National Laboratory, Livermore, California, USA
3Engine Research Center, University of Wisconsin, Madison, Wisconsin, USA
The manuscript was accepted after revision for publication on 19 April 2005.
DOI: 10.1243/146808705X30503

Abstract: Modelling the premixed charge compression ignition (PCCI) engine requires a
balanced approach that captures both uid motion as well as low- and high-temperature fuel
oxidation. A fully integrated computational uid dynamics (CFD) and chemistry scheme (i.e.
detailed chemical kinetics solved in every cell of the CFD grid) would be the ideal PCCI modelling
approach, but is computationally very expensive. As a result, modelling assumptions are
required in order to develop tools that are computationally ecient, yet maintain an acceptable
degree of accuracy. Multi-zone models have been previously shown accurately to capture
geometry-dependent processes in homogeneous charge compression ignition (HCCI) engines.
In the presented work, KIVA-3V is fully coupled with a multi-zone model with detailed chemical
kinetics. Computational eciency is achieved by utilizing a low-resolution discretization to
solve detailed chemical kinetics in the multi-zone model compared with a relatively highresolution CFD solution. The multi-zone model communicates with KIVA-3V at each computational timestep, as in the ideal fully integrated case. The composition of the cells, however,
is mapped back and forth between KIVA-3V and the multi-zone model, introducing signicant
computational time savings. The methodology uses a novel re-mapping technique that can
account for both temperature and composition non-uniformities in the cylinder. Validation
cases were developed by solving the detailed chemistry in every cell of a KIVA-3V grid. The
new methodology shows very good agreement with the detailed solutions in terms of ignition
timing, burn duration, and emissions.
Keywords: HCCI, PCCI, computer modelling, combustion processes, computational uid
dynamics, chemical kinetics, emissions
1 INTRODUCTION
1.1 HCCI and PCCI
Homogeneous charge compression ignition (HCCI)
has emerged in the last couple of decades as a
promising alternative to the well-established technologies of diesel and spark-ignited (SI) engines.
HCCI has the potential to deliver diesel-like fuel
* Corresponding author: Mechanical Engineering, The University
of Michigan, 2032 Walter E Lay Automotive Laboratory, 1231 Beal
Avenue, Ann Arbor, MI 48109-2133, USA. email: ababajim@
umich.edu

JER02305 IMechE 2005

conversion eciency, accompanied by low nitrogen


oxides (NO ) and soot emissions, particularly at partx
load, where its counterparts suer signicantly [13].
Several technical issues must, however, be resolved
before the concept nds an application in production engines. HCCI engines suer from high
unburned hydrocarbon (HC) and carbon monoxide
(CO) emissions. High rates of heat release, and the
resulting steep pressure rise rates and high-peak
pressures, limit the HCCI operating range (low specic
power output). The main obstacle to date is the lack
of any direct means of controlling the ignition timing,
such as the spark or late injection event available in
current production engines [4].
Int. J. Engine Res. Vol. 6

498

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

The control issue has prompted the investigation of


various control strategies, including direct injection
(DI) [5] and variable valve actuation (VVA), the latter
as a means of retaining large levels of residual gases
in the cylinder [69]. Both of these strategies can
result in a mixture with signicant composition and
temperature stratication. During gas exchange and
compression the actual time for mixing of the fresh
air, fuel, and the residuals is limited and, as a consequence, there are potential inhomogeneities in
the mixture [10]. This composition stratication
may be desirable, because it can help in controlling
the phasing of combustion, extending the low load
operating range, and lengthening the burn duration
at higher loads [11, 12]. This partially stratied HCCI
can be better described with the less restrictive term
premixed charge combustion ignition (PCCI), which
will be the focus of the presented work.
1.2 Multi-zone modelling overview
It is widely accepted that HCCI is essentially controlled by chemical kinetics, with little direct eect
of turbulence. During the main heat release, the
chemical kinetic processes occur on such short time
scales that turbulence is too slow to inuence the
process. Although there is still some debate about
the inuence of turbulence on HCCI combustion
[1314], spectroscopic and imaging investigations of
HCCI verify that simultaneous multi-point ignition
occurs with no ame propagation [1518], supporting the hypothesis that heat release is dominated by
chemical kinetics. Recent modelling work by Sankaran
and Im [19] studied the inuence of dissipation rate
and mixture inhomogeneities on the auto-ignition of
mixtures on an opposed ow conguration, where a
premixed fuel/air stream mixed with hot exhaust
gases. It was found that, although dissipation rate
and mixture inhomogeneity can have a signicant
eect on auto-ignition at lower mixing rates and
higher pressures [conditions typically observed near
top dead centre (TDC) in an HCCI engine], chemical
kinetics dominate the auto-ignition process and
diusion has little eect on the propagation of the
reaction front.
In theory, accurate analysis of PCCI combustion
could be achieved by fully integrating a computational
uid dynamics (CFD) code with a detailed chemical
kinetics code, which would solve the chemistry in
each computational cell. Such a model would require
grids ne enough (on the order of 104 to 105 cells) to
resolve adequately the temperature distribution in
the cylinder. The combination of these large numbers
of cells with computationally intensive chemical
Int. J. Engine Res. Vol. 6

kinetics calculations makes such a model impractical


even for todays fastest computers. For now, CFD
calculations with detailed chemistry in each cell
are limited to small chemical mechanisms (~100
species) and coarse grids (~1000 cells). Recent
publications [20, 21] have reported good results
using this approach.
In order to obtain some of the resolution aorded
by CFD models and yet reduce the computational
time required by chemical kinetics calculations,
Aceves et al. introduced a sequential multi-zone
modelling approach [2224]. This multi-zone model
of HCCI combustion assumes a decoupling of the
turbulent mixing process and chemistry prior to and
during the main heat release. The approach uses a
CFD code, KIVA-3V [25], and a multi-zone chemical
kinetics solver, in a sequential fashion. KIVA-3V is
run over part of the engine cycle, typically from
intake valve closing (IVC) until a transition point
before top dead centre (BTDC). At this point, cells
with similar pressure and temperature histories are
grouped into a relatively small number of zones
(10100). The chemical kinetics solver is then applied
to this small number of zones, instead of the large
number of cells used in the CFD code, oering a
great advantage in computational time compared
with a full integration of a chemical kinetics code
and a CFD code. The multi-zone model has been
successful in the prediction of combustion process
parameters, such as peak cylinder pressure and burn
duration. HC emissions are predicted reasonably well,
but CO is typically underpredicted by an order of
magnitude.
The limitation of the sequential method is that
once the chemistry calculation begins, the detailed
information from the CFD code is lost and there is
no mixing between the zones. Flowers et al. [26]
modied the multi-zone model to include mixing
eects, by introducing a coupled CFD/multi-zone
model. Instead of a one-way mapping of the CFD temperature distribution onto the multi-zone chemical
kinetics solver, their method was modied to have
mapping back and forth throughout the cycle. The
advantage is that the uid mechanical processes
are still calculated on a high-resolution grid, while
the computationally intensive chemical kinetics
processes can be solved within a small number of
zones. The goal of the study was to investigate the
mixing of cold gases from the crevices and the
boundary layer with the hot gases in the core during
the post main heat release oxidation. It was found
that there is diusion of CO and fuel from the lowest
temperature regions into the hotter gases in the
core and that the fuel continues to react during the
JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

expansion stroke. In addition, the predictions of HC


and CO emissions were improved compared with the
sequential multi-zone model.
The models of Aceves and Flowers were applied to
closed-cycle applications, starting at IVC, and took
into account only the temperature distribution in the
cylinder, while the initial composition was homogeneous [2228]. Babajimopoulos et al. [29] extended
the sequential CFD/multi-zone model for open-cycle
applications and investigated the mixing of residual
gases with the fresh charge for two VVA strategies
(residual gas fraction ranging from 30 to 60 per cent).
This study showed both temperature and composition
stratication caused by incomplete mixing of hot
residual gasses with the fresh charge. In order to
address the composition inhomogeneity during the
chemistry calculation, the fuelO equivalence ratio
2
was used as an additional criterion for grouping
cells into zones. KIVA-3V was used to simulate the
gas exchange processes (exhaust and intake) and
compression up to a transition point, after which
the multi-zone code calculated the combustion and
expansion processes. The transition point was dened
as the time at which the temperature of the hottest
cell in the cylinder exceeded 950 K. For the initialization of the zones, both the distributions of temperature and fuelO equivalence ratio were taken into
2
consideration and a two-dimensional (2D) array of
zones, with dierent compositions and temperatures,
was generated. It was shown that, under certain
conditions, the eect of composition stratication
is signicant and cannot be captured by a multizone model using only temperature zones. As in the
original multi-zone model by Aceves et al., there was
no mixing of the zones after the transition and all
spatial information from KIVA-3V was lost. Later
work by Aceves et al. [30] demonstrated that the
multi-zone solution is sensitive to the transition time.
1.3 Objectives
The objective of this work is to build on existing work
and develop a new methodology for the modelling
of PCCI combustion in the framework of KIVA-3V
that fully accounts for mixing throughout the cycle.
The goal is simultaneously to reduce the computational time required by the solution of detailed
chemistry in each computational cell while maintaining accurate prediction of combustion parameters,
such as pressure, temperature, and emissions.
In the proposed methodology a multi-zone model
solving detailed chemical kinetics is fully coupled with
KIVA-3V. The model communicates with KIVA-3V at
each computational timestep, and the composition of
JER02305 IMechE 2005

499

the cells is mapped back and forth between KIVA-3V


and the multi-zone model. The zone initialization
takes into consideration both the temperature and
the composition of the cells.
The paper is arranged as follows. Section 2 presents the numerical simulations for the validation
of the proposed methodology. Denitions for the
equivalence ratio are introduced in section 3, which
will be used for the grouping of the cells into
temperatureequivalence ratio zones. In section 4
the coupling of the multi-zone model with KIVA-3V
is described and two methods of mapping the composition back onto the KIVA-3V cells are compared.
Section 5 presents the conclusions.

2 NUMERICAL SIMULATIONS FOR VALIDATION


The development of any new model is usually
accompanied by validation against experimental data.
This validation can be particularly cumbersome when
the new model addresses only a small component
of the overall system and relies on other submodels
for input. For example, when validating a multizone PCCI combustion model in the framework of a
CFD code, dierences between analytical and experimental results can then be caused by deciencies in
the multi-zone model, deciencies in the mixing and
injection models used in the CFD code, or even
incorrect initial and boundary conditions (e.g. initial
temperature, wall temperatures, etc.).
As previously discussed, a fully integrated uid
mechanicschemical kinetics code, which solves for
the detailed chemistry in each computational cell,
can yield accurate results for PCCI combustion, provided that all associated submodels perform equally
well. Since the proposed model is presented as a
computationally ecient alternative to the solution
of detailed chemistry in each computational cell, the
authors deemed it appropriate to validate the model
results against analytical results obtained from a fully
integrated uid mechanicschemical kinetics code,
keeping all other CFD submodels exactly the same.
While it is ultimately necessary to conduct validation
against experimental data, the comparison with the
fully integrated results provides a good estimate of
the accuracy of the multi-zone methodology. Future
work will focus on validating the model against
experimental data.
During the development of the model, the code was
exercised and validated with a number of idealized
problems. These problems were simple enough to
be solved with a fully integrated KIVA-3V/chemical
kinetics code within a reasonable time. The engine
Int. J. Engine Res. Vol. 6

500

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

parameters and run conditions are listed in Table 1.


The KIVA-3V grid used for the analysis was a 2D
axisymmetric grid with 10 200 cells at bottom
dead centre (Fig. 1). The fuel was methane and the
chemical mechanism used in the analysis was GRIMech 3.0 [31]. The equivalence ratio and the residual
gas fraction were varied and several initial fuelair
and residual gas fraction (RGF) distributions at IVC
were used. These distributions were linear and were
imposed as only a function of the cylinder radius and
not a function of the axial position along the cylinder.
The initial fuel and RGF distributions for one of
the solved cases are shown in Fig. 1. This is a case
with a global equivalence ratio of 0.4 and 35 per cent
RGF. The initial fuelO equivalence ratio distribution
2
is quite steep, with values ranging from 0 to 1.1.
This was one of the most extreme cases that were
examined and the results presented in the following
sections correspond to this sample case. Agreement
between the multi-zone and the fully integrated
Table 1 Engine parameters and operating
conditions
Parameter

Value

Engine speed
Compression ratio
Stroke
Bore
Connecting rod length
Displacement
Fuel
Chemical kinetic mechanism
Start of calculation (IVC)
Pressure at IVC
Temperature at IVC
Average equivalence ratio
Residual gas fraction

2007 r/min
10.5:1
13.5 cm
11.41 cm
21.6 cm
1378 cm3
Methane
GRI-Mech 3.0 [31]
155
3.18 bar
565 K
0.30.4
235%

solutions improved as the fuelO equivalence ratio


2
distribution was made less steep and as the RGF was
reduced.
The running time for a fully integrated solution on
a 2.8 GHz PC operating on Linux was around 30 h.
The corresponding running time for a multi-zone
solution was approximately 3.5 h. This is a signicant
time gain, which is not accompanied by major loss
in accuracy, as will be demonstrated in the following
sections. It should be noted that the multi-zone
model would have a considerably greater advantage
in running time for higher-resolution grids. The
time for running the multi-zone model is largely
independent of the grid resolution, since chemistry,
which is the most time-consuming part of the calculation, is always solved for the same number of zones
(~100). On the other hand, the running time for the
fully integrated solution increases rapidly with the
number of KIVA-3V cells, since chemistry is solved
for every cell in the grid.

3 DEFINITIONS OF EQUIVALENCE RATIO FOR


COMPOSITION MAPPING AND REMAPPING
A key to the new model is nding an appropriate
composition marker to guide the correct mapping
and remapping of the cells composition between
KIVA-3V and the multi-zone model. In previous work
by the authors present [29, 30], when the multi-zone
code was used sequentially with KIVA-3V, the grouping of the cells into temperature (T )equivalence
ratio (W) zones took place only once during the cycle,
before any reactions had occurred, and thus the
fuelO equivalence ratio was an obvious choice. In
2

Fig. 1 Grid and sample distributions of CH and EGR (represented by CO ) at IVC for validation
4
2
runs
Int. J. Engine Res. Vol. 6

JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

the proposed methodology, however, the equivalence


ratio must provide more information about the composition and how combustion has progressed in the
cell. Before proceeding with the formulation of the
model, some denitions of equivalence ratio that will
be used in the following sections are introduced.
HCCI applications concern mainly lean or near
stoichiometric combustion with high levels of residual
gases. For W1 and no RGF, the complete combustion of a general oxygenated fuel of average
molecular composition C H O with air can be
x y z
written as

y z
WC H O + x+
(O +3.76N )
x y z
2
2
4 2
y
 xWCO + WH O
2
2 2

(1)

In equation (1), W is the traditionally dened


fuelair equivalence ratio [32]
(fuel/air)
a
(2)
(fuel/air)
s
where the subscript a denotes the actual fuelair
ratio and the subscript s the stoichiometric ratio.
When operating lean, the residual gas may contain
a signicant amount of excess oxygen. With high
levels of RGF a more appropriate equivalence ratio
is the fuelO equivalence ratio, W*
2
(fuel/O )
2 a
(3)
W*=
(fuel/O )
2 s
Equations (1) to (3) only relate to the composition
of the reactant and product species and do not
take into account the complex process by which
combustion proceeds. At any stage during the combustion process, the molecular composition is the
result of a global reaction of the form
W=

y z
X WC H O + x+
(O +3.76N )
x y z
2
2
4 2

K
 x C H O N
(4)
k ck hk ok nk
k=1
where K is the total number of species included in
the reaction mechanism, x is the molecular fraction
k
of species k in the mixture, and X is a normalizing
factor such that
K
x =1
k
k=1
JER02305 IMechE 2005

C H O N is a generic formula for species k,


ck hk ok nk
where
c , h , o , and n are the numbers of atoms
k k k
k
(carbon, hydrogen, oxygen, and nitrogen, respectively) that constitute the molecule of species k.
Knowing the mixture composition at each point
during the combustion process, the total number
of C atoms in the mixture, C#, is
K
K
K
C#= N c = x N c =N x c
(6)
k k
k tot k
tot
k k
k=1
k=1
k=1
where N is the number of molecules of species k
k
and N is the total number of molecules in the
tot
mixture. Similar expressions can be written for the
total number of atoms of oxygen (O#) and hydrogen
(H#). After some manipulation, an equation for the
global equivalence ratio, W, can be derived
W=

y z
[(1W)O +3.76N ]
+ x+
2
2
4 2

(5)

501

2C#+H#/2zC#
O#zC#

(7)

where z=z/x. The symbol used for the global


equivalence ratio is the same as for the traditional
fuelair equivalence ratio [equation (2)]. The reason
is that, for a fuelair mixture that has not reacted,
the two denitions are equivalent. The fundamental
dierence between the two denitions is that the
global equivalence ratio takes into account the
instantaneous atomic composition of the mixture,
while the traditional denition is based on the mass
fractions of the species. In a closed reactor, since the
mass and the total number of atoms of each element
are conserved, the global equivalence ratio is constant and is independent of whether combustion has
not started (reactants only), is already completed
(products only), or is at some intermediate stage.
It should be noted that the calculation of W with
equation (7) resembles the method used to estimate
the operating equivalence ratio of an engine from
the measured exhaust gas composition [32]. In
addition, the global equivalence ratio W as dened
in equation (7), is similar to the conserved scalar Z*
that Bilger uses in [33] to study the structure of nonpremixed ames.
An alternative denition for the equivalence ratio
(referred to as progress equivalence ratio, Q) can be
obtained if the number of C, H, and O atoms that
belong to the complete combustion products CO
2
and H O are excluded from the calculation in
2
equation (7)
2C#
+(H#
/2)zC#
CO2
H2O
CO2
(8)
O#
zC#
CO2H2O
CO2
W is directly related to the oxygen ratio V introduced
by Mueller et al. in [34]. Mueller et al. used the oxygen
ratio to measure the instantaneous stoichiometry of
Q=

Int. J. Engine Res. Vol. 6

502

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

a mixture accounting for the number of C, H, and O


atoms that belonged to fuel or oxidizer, and fuelplus-oxidizer reactant species.
For a closed reactor, the value of the progress
equivalence ratio can range from W (reactants only)
to 0 (complete combustion products), depending
on the progress of combustion. The dierences
between the fuelO equivalence ratio (W*), the global
2
equivalence ratio (W), and the progress equivalence
ratio (Q) are illustrated in Fig. 2. Figure 2 shows
the evolution of temperature and the equivalence
ratios during the combustion of four dierent fuels
(iso-octane [35], propane [36], methane [31], and
dimethyl ether (DME) [37]). The calculations were
performed with an adiabatic variable volume reactor
code, based on CHEMKIN [38].
The initial global equivalence ratio for all cases
was W=0.4 and the RGF, including only complete
combustion products, was 5 per cent. Calculations

started at 90 BTDC, the initial pressure was 1.8 bar,


and the initial temperature was adjusted so that the
main heat release occurred after TDC (ATDC). As
already mentioned, the global equivalence ratio is
constant throughout the whole process. The initial
value of the fuelO and the progress equivalence
2
ratio is the same and slightly lower than W. This is
attributed to the 5 per cent RGF that includes excess
O , which lowers the value of W* compared with W,
2
and CO and H O, whose atoms are not used in the
2
2
calculation of Q. W* and Q are equivalent, when there
are no intermediate combustion products present.
As it can be seen in Fig. 2, the value of W* starts
dropping well before the main heat release event,
as the fuel breaks down into smaller-chain hydrocarbons and radicals. This is more evident for isooctane, which is the largest molecule, and less for
methane, which has only one carbon atom. On the
other hand, the evolution of Q corresponds much

Fig. 2 Equivalence ratio and temperature evolution during combustion for various fuels, as
calculated using an adiabatic variable volume reactor simulation (the numbers in the
square brackets are the references for the chemical mechanisms used). The initial global
equivalence ratio for all cases was W=0.4 and the residual gas fraction was 5 per cent
Int. J. Engine Res. Vol. 6

JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

503

better with the heat release and the subsequent rise


of the temperature in the cylinder for all fuels than
W* does, owing to the fact that the main heat release
occurs when CO and H O are formed. It is, there2
2
fore, apparent that the progress equivalence ratio, Q,
is the most appropriate choice for identifying cells
with similar composition and it will be used in the
following sections for grouping cells into TQ zones.

4 MODEL FORMULATION
4.1 Coupled KIVA-3V and multi-zone model
KIVA-3V [25] provides the CFD framework for the
proposed methodology. KIVA-3V handles the uid
mechanical processes on a highly resolved grid, while
the computationally intensive chemical kinetics are
solved only for a relatively small number of zones.
The equations pertinent to chemical reactions that
need to be addressed by the multi-zone code are the
continuity and energy equations. The continuity
equation is

C A BD

qr
r
k +V(r u)=V rD V k
k
k
qt
r

+r c +r sd
k
kl

(9)

where r is the mass density of species k, r the total


k
mass density, and u the uid velocity vector. Ficks
law is used for diusion with a single coecient D .
k
The terms r c and r s indicate source terms resulting
k
from chemistry and spray, respectively. Species l is
the species of which the spray droplets are composed
and d is the Dirac delta function.
kl
The internal energy equation is
q(rI )
+V(ruI )=rVu+(1A )s : VuVJ
o
qt
c+Q
s
+A re+Q
(10)
o
where I is the specic internal energy exclusive of
chemical energy, p is the pressure, s : Vu indicates
the double-dot product between the surface tension
and velocity gradient tensor, J is the sum of the contributions owing to heat conduction and enthalpy
diusion, e is the dissipation rate of the turbulent
c and Q
s are source terms caused
kinetic energy, and Q
by chemical heat release and spray interaction.
What KIVA-3V requires from the chemistry solver
at each timestep is the composition change (r c )
kc)
and the heat release due to chemical reactions (Q
for each cell. Figure 3 shows schematically the procedure for the application of the current multi-zone
model to the calculation of PCCI combustion. At
each discrete time, t, every cell in KIVA-3V is at
JER02305 IMechE 2005

Fig. 3 Procedure for mapping between KIVA-3V and


multi-zone chemical kinetics solver

some thermodynamic state (temperature, pressure,


and composition). These cells are grouped into zones,
according to an algorithm that considers the temperature and progress equivalence ratio of the cells
and is presented in the following section. Each zone
contains a fraction of the mass in the cylinder. Thus,
each zone becomes representative of a group of cells
throughout the cylinder. The averages of temperature, pressure, and composition for the cells in a
zone are determined to specify the thermodynamic
state of the mixture in that zone.
The chemical process for each zone is handled by
CHEMKIN [38]. Each zone is allowed to react from
time t to time t+Dt in an adiabatic constant volume
reactor. The adiabatic reactor is used only to determine the change in composition and an amount of
heat release for each zone. The constant volume
assumption may introduce some error, but the timesteps used (on the order of 1 ms) are suciently
small to make it negligible [24]. Moreover, heat
transfer, convection, and diusion between the cells
(and thus the zones) are calculated by KIVA-3V. After
the chemistry calculation, the new zone composition
and the heat release are mapped back onto all the
cells within the zone. As will be shown in the next
Int. J. Engine Res. Vol. 6

504

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

sections, the way this remapping is performed is


critical for the quality of the solution. The KIVA-3V
code subsequently proceeds to determine convection
and diusion processes over the same timestep.

Table 2 Mass fractions in temperature zones


Zone (1=coldest, 5=hottest)

Mass fraction in zone, %

10

20

30

35

4.2 Grouping of cells into TQ zones


At each timestep, the grouping of the KIVA-3V
cells into zones must be performed in a way
which guarantees that the grouped cells have similar
thermodynamic and chemical properties. Temperature is the most representative variable of the
thermodynamic state of a cell. In addition, as already
shown in section 3, the progress equivalence ratio Q
provides information about the composition and
how far combustion has progressed in the cell. These
two variables are, therefore, selected and the cells are
grouped into TQ zones.
In previous work by the present authors [27] it was
shown that, when there is signicant temperature
and composition stratication in the cylinder, the
part of the charge that ignites and burns fastest is
not necessarily the hottest one, but rather one that
has a combination of relatively high temperature and
fuelO equivalence ratio. Temperature, however, is
2
still the dominant factor that determines ignition
timing and the equivalence ratio plays a secondary
role, especially for the high-octane-number fuels
typically used in PCCI engines. Taking this into consideration, the steps for grouping the cells into zones
are as follows.
1. All cells in the cylinder are sorted from lowest to
highest temperature and are divided into ve temperature zones. Each zone contains a prescribed
fraction of the mass within the cylinder, given in
Table 2.

2. The cells in each temperature zone are sorted from


lowest to highest Q. Starting from the cell with
the lowest Q, the cells of the temperature zone are
divided into as many zones as needed, so that the
maximum Q range in each zone is DQ =0.02.
max
Figure 4 shows an illustrative example of TQ zone
generation for the sample case presented in Fig. 1.
The left side of Fig. 4 shows the temperature and
the progress equivalence ratio distributions in the
cylinder at 20 BTDC, as obtained from KIVA-3V.
The right side of Fig. 4 shows the cells that belong
in the third temperature zone, divided into Q
zones.
3. The last step is to take any TQ zones that contain
more than 1 per cent of the cylinder mass, sort
the cells in these zones by temperature, and divide
them into smaller temperature zones so that, in
the end, the mass fraction in each zone does not
exceed 1 per cent.
This zone generation process yields a total number
of zones just over 100 (the mass fraction in some
zones can actually be less than 1 per cent). Once the
averages of temperature, pressure, and composition
for the cells in a zone are determined, the zone is
allowed to react from time t to time t+Dt. After the
chemistry calculation, the change in the zone composition and the energy released owing to chemical
reactions are mapped back onto the cells within the
zone. The next step is to remap the species from the

Fig. 4 Schematic of TQ zone generation. The temperature and progress equivalence ratio elds
are used to divide the cells rst into T zones, which are then divided into smaller Q zones.
The gure on the right shows the geometric shape of the third temperature zone, which
contains 20 per cent of the mass (between 15 and 35 per cent of the mass when ranked
according to temperature)
Int. J. Engine Res. Vol. 6

JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

chemical kinetic zones onto the KIVA-3V cells. Two


approaches to handling remapping will be discussed
in the following sections.
4.3 Remapping using average values
The simplest procedure to map zones back onto the
cells consists of assigning the average composition
of the zone to each cell. This means that all the cells
in a particular zone will have the same composition.
The energy per unit mass released owing to chemical
reactions, De, as computed by the chemistry calculation, is mapped onto the cells in the zone as a
change in the specic internal energy. The new cell
temperature is then computed using the updated
specic internal energy and the remapped average
composition. This simple remapping process is
referred to as average remapping.
Figure 5 compares the solution obtained with the
multi-zone model with average remapping against
the detailed solution with full chemistry in each cell
for the sample case shown in Fig. 1. The variables
compared are cylinder pressure, mean temperature,
and maximum temperature in the cylinder. The
cylinder pressure and mean temperature are in
good agreement, but the maximum temperature is
underpredicted with the multi-zone model.
In order to understand the discrepancy in the
maximum temperature prediction, it is useful to look
at what happens in the temperature and composition
distributions in the cylinder during combustion.
A way to do this is to plot the TW distribution
in the cylinder at dierent times. These plots are
essentially scatter plots of the temperature and global
equivalence ratio of all the computational cells in the
cylinder and resemble clouds [29]. The linear pattern
of these plots is a result of heat transfer and the temperature distribution in the cylinder. The layer of
cells closest to the cylinder wall is the coldest, while
the cells closer to the centre of the cylinder have a
more uniform temperature (see Fig. 4). As a result,
the low-temperature cells are further apart than

505

the higher-temperature cells. This pattern would be


less prominent if a higher-resolution grid had been
used. Figure 6 compares the TW cloud evolution
for the multi-zone model with average remapping
(black points) against the detailed solution (grey
points) at four times: the transition point at which
the multi-zone model is called for the rst time
and is dened as the time at which the temperature
of the hottest cell in the cylinder exceeds 900 K
(~41 BTDC), ve timesteps after the transition
point (Dt=1 ms), 40 BTDC, and TDC.
The eect of grouping into zones and remapping
the average zone composition onto the cells is
evident. Early in the process bands of cells with
the same global equivalence ratio are formed. This
introduces large articial composition gradients
in KIVA-3V, which signicantly enhance diusion
between neighbouring cells that belong to dierent
zones [equation (9)]. Remapping the average zone
composition in each cell essentially increases the
eective grid size from the dimension of one cell to
the dimension of one zone, thus introducing large
numerical diusion. Only 1 crank angle degree (CAD)
after the transition point, the shape of the TW cloud
for the multi-zone has changed signicantly and its
edges have disappeared. By TDC, the cloud has
almost collapsed to a straight line and all the cells in
the cylinder have the same global equivalence ratio.
The temperature distribution, however, has been
maintained to a large degree, because the energy
released owing to chemical reactions is mapped onto
the cells in the zone as a change in the specic
internal energy, thus preserving the original temperature distribution throughout the zone. This means
that the multi-zone model with average remapping
is eectively solving a dierent problem with all cells
having the same average composition rather than the
initially specied distribution. This also explains why
the maximum temperature in the cylinder is underpredicted, since the combustion temperature is lower
for lower equivalence ratios.

Fig. 5 Comparison of simulation results of multi-zone solution with average remapping against
the detailed solution
JER02305 IMechE 2005

Int. J. Engine Res. Vol. 6

506

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

Fig. 6 TW cloud evolution for the detailed solution and the multi-zone solution with average
remapping

From the discussion above, it is apparent that


average remapping does not fully capture the physical
processes and an alternative remapping method
needs to be developed.
4.4 Improved remapping to maintain
composition and temperature gradients
Once cells are grouped into a zone, and the zone
is permitted to react using the average temperature
and composition of the cells, it is impossible to
know exactly what fraction of the mass of each
species should be mapped back onto each cell. This
knowledge can be obtained only by solving the
chemistry in each single cell (fully integrated solution). It is, however, possible to distribute the species
back to the cells, so that the thermodynamic properties of the cells do not change signicantly. In order
to do that, certain requirements have been specied.
1. The mass of each cell in the zone must be conserved.
2. The mass of each individual species in the zone
must be conserved.
Int. J. Engine Res. Vol. 6

3. The number of C, H, O, and N atoms in each cell


must be conserved.
The last requirement is the most dicult to achieve,
so instead it is attempted to keep the number of C, H,
O, and N atoms in each cell as close to the original
value (before the chemistry calculation) as possible.
The steps of the new remapping algorithm (referred
to as improved remapping) are outlined below.
A new quantity, ch, is dened based on the number
of C and H atoms included in all the species except
CO and H O
2
2
H#
(11)
ch=2C#
+ H2O
CO2
2
This new quantity is similar to the numerator of the
progress equivalence ratio [equation (8)] and is calculated for the zone (ch
) and each individual cell
zone
(ch ) before the chemistry calculation. The sum of
cell
ch
of all individual cells in a particular zone is
cell
equal to the zones ch
.
zone
After the chemistry calculation, all the species,
except CO , H O, O and N , are distributed to the
2
2
2,
2
JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

zones cells based on ch. The mass of species k in


each cell will be
ch
cell m
(12)
k,zone
ch
zone
Equation (12) guarantees that the mass of each
individual species in the zone is conserved. This
interpolation may actually increase slightly the
number of C or H atoms in some cells. If this
happens, then the total number of C or H atoms that
the remaining cells in the zone are allowed to have
is adjusted accordingly, so that the total number of
m

k,cell

507

C and H atoms in the zone remains constant. For the


most part, however, there is still a shortage of C and
H atoms in each cell. The only remaining species
containing C is CO , which is distributed to the zones
2
cells, lling up the missing C atoms (M is the
k
molecular mass of species k)
m
m
k,cell c + CO2,cell =C#
(13)
k
cell
M
M
k
CO2
k
Similarly to CO , H O can be distributed to the
2
2
zones cells to maintain the number of H atoms in
each cell. Having done this the only species that have

Fig. 7 Comparison of simulation results of multi-zone solution with improved remapping


against the detailed solution

Fig. 8 TW cloud evolution for the detailed solution and the multi-zone solution with improved
remapping
JER02305 IMechE 2005

Int. J. Engine Res. Vol. 6

508

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

not been distributed are O and N . O is distributed


2
2 2
to maintain the total number of O atoms in each
cell. Finally, N is used as a ller, to bring each cell
2
to its original mass. In the last step, the change
in the specic internal energy of each cell is calculated accounting for the internal energy of formation
of the species present in the cell before grouping
the cells into zones and after the chemistry calculation and the remapping have been performed. The
new cell temperature is then computed using the
updated specic internal energy and the remapped
composition.
Figure 7 compares the solution obtained using the
multi-zone model with the improved remapping
against the detailed solution with full chemistry in
each cell for the sample case shown in Fig. 1. The
agreement between the two solutions for pressure
and mean temperature is excellent and the maximum temperature prediction has improved as well,
compared with the average remapping.
The benets of the improved remapping can be
seen more clearly if the TW and TQ clouds for the

multi-zone model with improved remapping and


the detailed solution are compared at dierent times
in Figs 8 and 9. Figure 8 compares the TW clouds at
40 BTDC, TDC, 5 ATDC and 15 ATDC. Up to TDC,
which is before any signicant heat release has
occurred, the two clouds are identical. At 5 ATDC,
the heat release has just started for the detailed
solution, with the multi-zone solution following
closely. As already mentioned, it is the cells with high
temperature and equivalence ratio (top-right corner
of the cloud) that burn rst. At 15 ATDC, the main
heat release event is over and the TW cloud of the
multi-zone code with the improved remapping has
caught up and is very similar to the detailed solution.
The eect that combustion has had on the shape of
the TW cloud is raising the temperature of the cells,
moving the cloud along the T axis. The composition
of the cells has of course changed, but the global
equivalence ratio remains the same as reactants are
converted to combustion products.
Figure 9 compares the TQ clouds at TDC, 5 ATDC,
and 15 ATDC. As with the TW clouds, up to TDC

Fig. 9 TQ cloud evolution for the detailed solution and the multi-zone solution with improved
remapping
Int. J. Engine Res. Vol. 6

JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

there are no dierences between the multi-zone


solution with improved remapping and the detailed
solution. At 5 ATDC, the shape of the cloud for the
detailed solution starts changing signicantly as the
progress equivalence ratio of the cells that have
started burning goes to 0 and their temperature
increases. After the end of the main heat release
event, at 15 ATDC, the progress equivalence ratio of
most cells is almost 0. There are, however, still cells
with high Q and relatively low temperature. These

509

cells have not burned completely and are the source


of unburned HC and CO emissions.
So far, the two remapping methods the average
and the improved have been compared with the
detailed solution only in terms of pressure and
temperature predictions and TW cloud evolution.
Figure 10 compares the evolution of the mass
fractions of six species in the cylinder (CH , CO ,
4
2
CO, NO, OH, and H O ) for the two remapping
2 2
methods with the detailed solution. The improved

Fig. 10 Comparison of the evolution of the mass fraction of selected species using the multizone model with the two dierent remapping options (average and improved) against
the detailed solution
JER02305 IMechE 2005

Int. J. Engine Res. Vol. 6

510

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

remapping performs signicantly better than the


average remapping for all species, especially later
in the expansion stroke (50 ATDC), when the composition in the cylinder begins to become frozen.
More specically, the average remapping method
overpredicts the formation of CO (combustion
2
eciency), while it underpredicts the nal CO mass
fraction. In addition, the average remapping signicantly underpredicts the formation of NO, which is
directly related to the underestimated in-cylinder
peak temperature. Overall, the improved remapping
does an excellent job in tracking down the evolution
of all the species in the cylinder. Any large discrepancy with the detailed solution is short lived and only
during the period of high heat release rate, when
temperature and composition change rapidly in each
cell. By the end of combustion, however, the nal
predictions of the improved remapping method are
very close to the detailed solution.

5 CONCLUSIONS
A multi-zone model with detailed chemical kinetics
has been fully coupled with KIVA-3V for the simulation of PCCI engines. The main conclusions of this
study are outlined below.
1. The computational time with the multi-zone model
reduced the simulation time by almost 90 per cent
(from 30 to 3.5 h) relative to solving chemistry
directly in every cell of the KIVA-3V grid. The grid
used for validation had a relatively small number
of cells compared with what would typically
be used for engine simulations. The multi-zone
solution of chemistry uses essentially the same
number of zones to solve chemistry independent
of the number of cells in the KIVA-3V grid. Thus,
the time benet of the multi-zone model relative
to solving chemistry in every cell would be much
greater for larger grids.
2. The traditional denitions of equivalence ratio
fuelair and fuelO are not sucient when
2
trying to track the composition in each cell
throughout the combustion process. For this
reason, two new denitions of equivalence ratio
were introduced, global (W) and progress (Q);
these are calculated using the total number of C,
H, and O atoms in each cell. It was shown that
the evolution of the progress equivalence ratio
corresponds well with heat release. The progress
equivalence ratio is, therefore, used for grouping
cells into TQ zones.
Int. J. Engine Res. Vol. 6

3. Two methods for mapping the new zone composition back onto the KIVA-3V cells were examined. In the rst approach (average remapping),
the average composition of the zones is mapped
onto the cells of each zone. It was found that
this approach introduces signicant articial
composition gradients in the cylinder resulting
in non-physical increased diusion and mixing.
The second remapping approach (improved
remapping) uses an algorithm that attempts to
maintain the total number of C, H, O, and N
atoms in each cell constant. It was shown that the
non-physical diusion is avoided with improved
remapping and predictions of pressure, temperature, and species mass fractions match very
closely the solution of chemistry in every cell.
4. Overall, the proposed methodology oers a computationally ecient alternative to the CFD with
detailed chemistry in every cell approach, while
maintaining good agreement with the detailed
solution. The fully coupled KIVA-3V and multizone model can provide a useful tool for the
fundamental understanding of PCCI combustion,
the eects of temperature and composition stratication on ignition and burn duration, and the
sources of emissions.

REFERENCES
1 Najt, P. M. and Foster, D. E. Compression-ignited
homogeneous charge combustion. SAE paper
830264, 1983.
2 Thring, R. H. Homogeneous-charge compressionignition (HCCI) engines. SAE paper 892068, 1989.
3 Christensen, M., Johansson, B., and Einewall, P.
Homogeneous charge compression ignition (HCCI)
using isooctane, ethanol and natural gas a comparison with spark ignition operation. SAE paper
972874, 1997.
4 Willand, J., Nieberding, R.-G., Vent, G., and
Enderle, C. The knocking syndrome: its cure and
potential. SAE paper 982483, 1998.
5 Marriot, C. D. and Reitz, R. D. Experimental investigation of direct injection-gasoline for premixed
compression ignited combustion phasing control.
SAE paper 2002-01-0418, 2002.
6 Kontarakis, G., Collings, N., and Ma, T. Demonstration of HCCI using a single-cylinder, four-stroke
SI engine with modied valve timing. SAE paper
2000-01-2870, 2000.
7 Kaahaaina, N. B., Simon, A. J., Caton, P. A., and
Edwards, C. F. Use of dynamic valving to achieve
residual-aected combustion. SAE paper 2001-010549, 2001.
JER02305 IMechE 2005

A fully coupled CFD and multi-zone model

8 Law, D., Kemp, D., Allen, J., Kirkpatrick, G., and


Copland, T. Controlled combustion in an IC-engine
with a fully variable valve train. SAE paper 2001-010251, 2001.
9 Koopmans, L. and Denbratt, I. A four stroke camless
engine, operated in homogeneous charge compression ignition mode with commercial gasoline.
SAE paper 2001-01-3610, 2001.
10 Babajimopoulos, A., Assanis, D. N., and Fiveland,
S. B. An approach for modeling the eects of gas
exchange processes on HCCI combustion and its
application in evaluating variable valve timing control strategies. SAE paper 2002-01-2829, 2002.
11 Dec, J. E. and Sjo
berg, M. A parametric study of
HCCI combustion the sources of emissions at low
loads and the eects of GDI fuel injection. SAE
paper 2003-01-0752, 2003.
12 Wolters, P., Salber, P., Geiger, J., Duesmann, M.,
and Dilthey, J. Controlled auto ignition combustion
process with an electromechanical valve train. SAE
paper 2003-01-0032, 2003.
13 Kong, S.-C., Marriot, C. D., Reitz, R. D., and
Christensen, M. Modeling and experiments of HCCI
engine combustion using detailed chemical kinetics
with multidimensional CFD. SAE paper 2001-011026, 2001.
14 Christensen, M. and Johansson, B. The eect of
in-cylinder ow and turbulence on HCCI operation.
SAE paper 2002-01-2864, 2002.
15 Onishi, S., Jo, S. H., Shoda, K., Jo, P. D., and Kato, S.
Active thermo-atmosphere combustion (ATAC) a
new combustion process for internal combustion
engines. SAE paper 790501, 1979.
16 Noguchi, M., Tanaka, Y., Tanaka, T., and
Takeuchi, Y. A study on gasoline engine combustion
by observation of intermediate reactive products
during combustion. SAE paper 790840, 1979.
17 Iida, N. Combustion analysis of methanol-fueled
active thermo-atmosphere combustion (ATAC)
engine using a spectroscopic observation. SAE
paper 940684, 1994.
18 Hultqvist, A., Christensen, M., Johansson, B.,
Richter, M., Nygren, J., Hult, J., and Alden, M.
The HCCI combustion process in a single cycle
high-speed fuel tracer LIF and chemiluminescence
imaging. SAE paper 2002-01-0424, 2002.
19 Sankaran, R. and Im, H. G. Eects of mixture
inhomogeneity on the auto-ignition of reactants
under HCCI environment. Proceedings of the 42nd
AIAA Aerospace Sciences Meeting and Exhibit, 2004,
AIAA Paper 2004-1328.
20 Kusaka, J., Tsuziki, K., Daisho, Y., and Saito, T. A
numerical study on combustion and exhaust gas
emissions characteristics of a dual-fuel natural gas
engine using a multi-dimensional model combined
with detailed kinetics. SAE paper 2002-01-1750,
2002.
21 Kong, S.-C., Reitz, R. D., Christensen, M., and
Johansson, B. Modelling the eects of geometrygenerated turbulence on HCCI engine combustion.
SAE paper 2003-01-1088, 2003.
JER02305 IMechE 2005

511

22 Aceves, S. M., Flowers, D. L., Westbrook, C. K.,


Smith, J. R., Pitz, W., Dibble, R., Christensen, M.,
and Johansson, B. A multi-zone model for prediction of HCCI combustion and emissions. SAE
paper 2000-01-0327, 2000.
23 Aceves, S. M., Flowers, D. L., Martinez-Frias, J.,
Smith, J. R., Westbrook, C. K., Pitz, W. J., Dibble, R.,
Wright, J. F., Akinyemi, W. C., and Hessel, R. P. A
sequential uidmechanic chemicalkinetic model
of propane HCCI combustion. SAE paper 2001-011027, 2001.
24 Aceves, S. M., Martinez-Frias, J., Flowers, D. L.,
Smith, J. R., Dibble, R. W., Wright, J. F., and
Hessel, R. P. A decoupled model of detailed uid
mechanics followed by detailed chemical kinetics
for prediction of iso-octane HCCI combustion. SAE
paper 2001-01-3612, 2001.
25 Amsden, A. A. KIVA-3V: A block-structured KIVA
program for engines with vertical or canted valves.
Los Alamos National Laboratory Report LA-13313MS, 1997.
26 Flowers, D., Aceves, S., Martinez-Frias, J., Hessel, R.,
and Dibble, R. Eect of mixing on hydrocarbon
and carbon monoxide emissions prediction for isooctane HCCI engine combustion using a multi-zone
detailed kinetics solver. SAE paper 2003-01-1821,
2003.
27 Flowers, D., Aceves, S., Smith, R., Torres, J.,
Girard, J., and Dibble, R. HCCI in a CFR engine:
experiments and detailed kinetic modeling. SAE
paper 2000-01-0328, 2000.
28 Aceves, S. M., Flowers, D. L., Espinosa-Loza, F.,
Martinez-Frias, J., Dibble, R. W., Christensen, M.,
Johansson, B., and Hessel, R. P. Piston-liner crevice
geometry eect on HCCI combustion by multi-zone
analysis. SAE paper 2002-01-2869, 2002.
29 Babajimopoulos, A., Lavoie, G. A., and Assanis, D. N.
Modeling HCCI combustion with high levels of
residual gas fraction a comparison of two VVA
strategies. SAE paper 2003-01-3220, 2003.
30 Aceves, S. M., Flowers, D. L., Espinosa-Loza, F.,
Babajimopoulos, A., and Assanis, D. N. Analysis
of premixed charge compression ignition combustion with a sequential uid mechanicsmultizone
chemical kinetics model. SAE paper 2005-01-0115,
2005.
31 Smith, G. P., Golden, D. M., Frenklach, M.,
Moriarty, N. W., Eiteneer, B., Goldenberg, M.,
Bowman, C. T., Hanson, R. K., Song, S.,
Gardiner, W. C., Lissianski, V. V., and Qin, Z.
http://www.me.berkeley.edu/gri_mech/
32 Heywood, J. B. Internal Combustion Engine
Fundamentals, 1988 (McGraw-Hill, New York).
33 Bilger, R. W. The structure of turbulent nonpremixed ames. Proc. Combust. Inst., 1988, 22,
475488.
34 Mueller, C. J., Pickett, L. M., Siebers, D. L., Pitz, W. J.,
Westbrook, C. K., and Martin, G. C. Eects of
oxygenates on soot processes in DI diesel engines:
experiments and numerical simulations. SAE paper
2003-01-1791, 2003.
Int. J. Engine Res. Vol. 6

512

A Babajimopoulos, D N Assanis, D L Flowers, S M Aceves, and R P Hessel

35 Curran, H. J., Gauri, P., Pitz, W. J., and Westbrook,


C. K. A comprehensive modeling study of iso-octane
oxidation. Combust. Flame, 2002, 129, 253280.
36 Marinov, N. M., Pitz, W. J., Westbrook, C. K.,
Vinvitore, A. M., Castaldi, M. J., and Senkan, S. M.
Aromatic and polycyclic aromatic hydrocarbon
formation in a laminar premixed n-butane ame.
Combust. Flame, 1998, 113, 192213.

Int. J. Engine Res. Vol. 6

37 Fischer, S. L., Dryer, F. L., and Curran, H. J.


The reaction kinetics of dimethyl ether. I: hightemperature pyrolysis and oxidation in ow reactors.
Int. J. Chem. Kinet, 2000, 32, 713740.
38 Kee, R. J., Rupley, F. M., and Miller, J. A.
CHEMKIN-II: A fortran chemical kinetics package
for the analysis of gas-phase chemical kinetics.
Sandia Report SAND89-8009, 1989.

JER02305 IMechE 2005

You might also like