You are on page 1of 10

Energy Conversion and Management 50 (2009) 11921201

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Assessing the effect of mass transfer on the formation of HC and CO emissions


in HCCI engines, using a multi-zone model
N.P. Komninos *
School of Mechanical Engineering, National Technical University of Athens, 9 Heroon Polytechniou St., Zografou Campus, 15780 Athens, Greece

a r t i c l e

i n f o

Article history:
Received 29 June 2008
Accepted 27 January 2009
Available online 24 February 2009
Keywords:
HCCI
Multi-zone model
Mass transfer
Emissions formation
Hydrocarbons
CO

a b s t r a c t
The focus of the present study is to assess the effect of mass transfer on the formation of unburned HC
and CO emissions in HCCI engines. A multi-zone model was modied and used for this purpose. The
new feature of the multi-zone model is its ability to switch between two distinct simulation modes,
i.e. either including or excluding mass transfer between zones. The switch between modes occurs at a
user-dened point in the engine closed cycle. Apart from mass transfer, the two modes use identical
sub-models for the heat transfer between zones and to the cylinder wall and for combustion simulation,
which is modeled using a reduced set of chemical reactions coupled with a chemical kinetics solver.
Using the modied multi-zone model, four cases were simulated and compared: one including mass
transfer throughout the closed cycle, and three cases whereby mass transfer is neglected after the initiation of the 1st or 2nd heat release or after the completion of main heat release. The simulation results
reveal that mass transfer affects the HC and CO accumulated at the colder regions during combustion and
governs the HC partial oxidation and CO production during expansion. For the operating conditions studied, neglecting mass transfer during combustion results to an underprediction of HC by as much as 50%
and of CO by 45% relative to the case where mass transfer is considered for. Omitting mass transfer only
during expansion, results to an overestimation of HC by 9% and to an underestimation of CO by 26%.
2009 Elsevier Ltd. All rights reserved.

1. Introduction
Homogeneous Charge Compression Ignition (HCCI) engines
have the potential of high thermal efciency and low NOx and soot
emissions as compared to conventional CI or SI engines [1]. However, operational and environmental issues arise since ignition
and combustion cannot be directly controlled and moreover unburned HC and CO emissions remain at a high level [13]. It is believed that HC and CO are formed at relatively cold regions within
the combustion chamber, i.e. in the crevice regions and the thermal
boundary layer [48]. Modeling results also support this hypothesis, i.e. that CO results from the partial oxidation of fuel originating
from the crevices and the thermal boundary layer [9,10].
Several models have been constructed for the description of
HCCI combustion. Easley et al. [9] used six zones to describe the
physical processes within the engine cylinder; three adiabatic constant mass core zones, an outer core zone exchanging mass with
the fth zone, representing the boundary layer and a crevice zone.
The thermal boundary layer was assigned constant thickness. Mass
and heat exchange was not considered between all zones. The
combustion process was modeled using chemical kinetics and a
set of chemical reactions. The temperature and mass distribution
* Tel.: +30 210 7723651.
E-mail address: nkom@central.ntua.gr
0196-8904/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2009.01.026

Downloaded from http://www.elearnica.ir

inside the combustion chamber at IVC was predened. It was


shown that both the crevice regions and the boundary layer have
an impact on unburned HC and CO emissions.
The same structure was used by Ogink and Golovitchev [11].
The authors used nine zones and estimated the temperaturevolume fraction distribution at IVC and an estimation for the thermal
boundary layer thickness which is held constant throughout compression and expansion. They also did not include both heat and
mass transfer between all zones, which affects the distribution
and magnitude of CO and unburned HC emissions as well as the
combustion mechanism. Xu et al. [12] also used the scheme introduced by Easley et al., whilst neglecting heat and mass transfer between zones. The model used underpredicted CO emissions and
overpredicted HC emissions.
Aceves et al. [13,14] used the KIVA code to generate temperaturemass distributions during compression and up to ignition. A
10-zone model was then used to simulate the combustion processes. No interaction between zones, neither heat nor mass exchange, was allowed. In [13] both unburned HC and CO
emissions were underpredicted, whilst in [14] unburned HC emissions were overpredicted for most of the cases examined, whilst
CO emissions were underpredicted. The demonstration of the
importance of mixing on the HC and CO emissions formation was
realized in a later study by Flowers et al. [15] by coupling CFD code
and chemical kinetics.

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

1193

Nomenclature
ah, bh, ch
B
K
MW
n
q_
Ru
Rmf
r
S
T
t
u
V
Y
W
Z

heat transfer constants


cylinder bore (m)
thermal conductivity (W/mK)
average zone molecular weight
moles (kmol)
heat ux (W/m2)
universal gas constant 8.314 (kJ/kmol K)
fuel burning rate (kg/m3 s)
distance (m)
cylinder height (m)
temperature (K)
thickness (m), time (s)
velocity (m/s)
volume (m3)
mass fraction ()
molecular weight of species
number of zones

Greek symbols
constant
Karman constant
dynamic viscosity (kg/m s2)
density (kg/m3)
molar rate of production (moles/(cm3 s))

a
j
l
q
x_

Subscripts
cyl
cylinder
g
gas
l
laminar
min
minimum

The present study focuses on the effect of mass transfer on the


formation of HC and CO using a multi-zone model. The multi-zone
model used is a modication of a previously presented model [16
18]. The major modication is the ability to switch between two
distinct modes of simulation; one mode considers for mass transfer
between zones and the other neglects mass transfer. The switch
between the two modes can be realized at a user-dened point
of the closed cycle in terms of crank angle. Thus the effect of mass
transfer on the performance and emissions formation can be determined in a straightforward manner.
In both modes, the multi-zone model consists of zones, which
assume spatial locations and dimensions. One of the zones represents the crevice regions, since these regions apart from being a
source for unburned HC, decrease the effective compression ratio
and the available chemical energy during main combustion. At
the mode where mass transfer is considered for, the zones exchange mass to maintain the in-cylinder pressure uniform, taking
into account their spatial location. As regards heat transfer and
combustion, the two modes use identical sub-models, features of
which have been used by other researchers [1921]. Specically,
in the work of Jia et al. [19] the inuence of mass transfer has also
been pointed out using features of the sub-models presented
herein. The variation of mixture composition due to combustion
is determined using chemical kinetics and Chemkin v.4 libraries
[22].
Four cases where examined and compared; in the rst case
mass transfer is included for the entire closed cycle, i.e. throughout
compression combustion and expansion. In the second and third
case mass transfer is considered for up to the initiation of the 1st
and 2nd heat release respectively. Beyond this point, the switch
between modes is realized and the zones retain constant mass.
In the fourth case mass transfer is included up to the completion

n
prod
t
trans
tot
w

normal
produced
turbulent
transferred
total
wall

Superscripts

characteristic value
+
dimensionless value
Dimensionless numbers
Pr
Prandtl number
Re
Reynolds number
Abbreviations
CA
crank angle
CFD
computational uid dynamics
CI
compression ignition
CR
compression ratio
CZM
constant zone mass
EVO
exhaust valve opening
HC
hydrocarbons
HRR
heat release rate
IVC
inlet valve closing
SI
spark ignition
TDC
top dead center
WMT
with mass transfer

of the 2nd heat release, beyond which the constant zone mass
mode is applied.
The comparison of the aforementioned cases revealed that mass
transfer affects to some extent the temperature eld and peak
combustion pressure, due to the signicant enthalpy transfer
occurring between zones mainly during combustion. As regards
HC and CO emissions, mass transfer affects the HC and CO accumulated at the colder regions during combustion and governs the HC
partial oxidation and CO production during expansion. For the
operating conditions studied, neglecting mass transfer during combustion results to a severe HC and CO underprediction relative to
the case where mass transfer is considered for. Omitting mass
transfer only during expansion, results to a mild overestimation
of HC and to a signicant underestimation of CO.
Although the results presented herein where obtained from a
multi-zone model, they are qualitative valid. Data obtained from
CFD models or their hybrids could be very useful for further investigation and for the more accurate description of mass transfer inside the combustion chamber. CFD data can also be used to further
validate the multi-zone model presented herein. These issues are
the subject of future research.
2. Model description
The basic structure of the multi-zone model used in the present
study has been presented in the past [1618]. The major modication made herein is the ability to switch between two modes; the
mode which includes mass transfer between zones (with mass
transfer WMT mode) and the mode which neglects mass transfer,
assuming constant zone mass (CZM mode). The switch between
the two modes is realized at a user-dened crank angle (CA) in
the closed cycle. Apart from mass transfer, the two modes of the

1194

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

model use identical sub-models for the description of heat transfer


and combustion.
2.1. Zone conguration
The conguration of the zones in both modes is identical. As
shown in Fig. 1 the combustion chamber is divided into three different types of zones, i.e. the core zone, the outer zones and the
crevice zone. The core zone is a cylinder, each of the outer zones
is a cylindrical annulus and the crevice zone lies beneath the outmost zone. The crevice zone represents all the crevice regions
which communicate directly with the combustion chamber, i.e.
the region above and behind the rst compression ring, the head
gasket crevice, etc.
The difference between the two modes lies in the calculation of
zone volumes and thicknesses. For the WMT mode, the crevice volume is a fraction of the clearance volume, whilst the volume of the
remaining zones varies throughout the engine cycle and is calculated according to their thickness. The thickness of the core zone
equals half its height and therefore varies throughout the engine
cycle. The thickness of all outer zones is the same in the x and
y direction and is constant during the cycle. From the previous
it is concluded that at TDC the sum of zone thickness must equal
half the cylinder clearance height:



z
X

ti 

i2

TDC

Smin
2

For the CZM mode, in which mass transfer is neglected, the volume of each zone, including the crevices, is calculated from thermodynamic considerations. The thickness of the zones, excluding
crevices, is thereafter determined from the zone conguration
and the volume of each zone. The sum of all zone volumes must
equal the combustion chamber volume and therefore condition
(1) is superuous in this mode. It is obvious that in CZM mode
the thickness of the zones varies during the calculations and generally differs from the zone thickness of the WMT mode. The computational procedure is explained in detail in Section 2.6.
2.2. Heat transfer
In the present study a modication has been made for the calculation of the heat transfer rate to the cylinder wall, in order to
reduce the user dened constants, which where used in the previous versions of the model [1618]. In those versions an Annandlike correlation was used for the estimation of the heat ux to
the cylinder wall [23]:

q_ w ah  k 

Rebh
 T g  T w ch  T 4g  T 4w
B


@T 
wall boundary condition
@r n rn 0

ti
liner

y
x
1
B
Fig. 1. Geometric conguration of the multi-zone model.

where kw is the conductivity of the charge at the cylinder wall



@T 
temperature, rn is the normal distance from the wall and @r

n
rn 0

the temperature gradient at the wall. The temperature gradient at


the wall is approximated numerically by the relations:



2 
1
@T 
T2  Tw
T3  T2

;

@rn rn 0
t2 =2
t 2 t3 =2
and T w < T

@T 
T2  Tw

;
@rn rn 0
t2 =2

when T w < T 2 < T 3


4a

in all other cases

4 b

where T is the mean charge temperature. The use of Eq. (4a) for the
estimation of heat ux gives better results than the simpler (4b)
during the latter stage of compression, combustion and expansion,
since it takes into account the change of temperature gradient when
moving from zone 3 to zone 2 to the wall. It is shown subsequently
that these equations provide an adequate agreement between
experimental and calculated pressure traces, which sufces for
the present study. A further discussion of the wall heat transfer is
out of the scope of the investigation presented, since the main focus
is the effect of mass transfer on the formation of HCCI engine emissions. The determination of the wall heat ux of an HCCI engine is
an active area of research and a heat transfer model commonly accepted has not yet been presented.
In the multi-zone model, heat is also transferred between zones
with a mechanism similar to conduction, i.e. the heat ux between
neighbouring zones is based on their temperature difference and
mean distance (Eq. (5)):

q_ ktot

@T
@r n

The amount of heat exchanged is calculated by multiplying the


heat ux by the surface area that separates them, and the time step
used.
For the determination of the total conductivity in Eq. (5), the approach of Yang and Martin [25] is followed:

The Yang and Martin approach has also been used by other
authors [26] for the CFD modelling of an HCCI engine, for the estimation of wall heat transfer in premixed charge engine combustion [27] as well as in HCCI multi-zone models [1921]. The ratio
of turbulent to laminar conductivity is calculated using the following formula:

cylinder head

q_ w kw

ktot kl kt

ti

2 3 i z

where Tg is the temperature of the outmost zone, Tw the wall temperature and ah, ch and bh, were the user dened constants.
In the present study the no-slip condition has been applied instead [24]. According to the no-slip boundary condition, the uid in
contact with the combustion chamber wall assumes the velocity of
the boundary (wall) and is considered therefore stationary. Consequently, heat is transferred through this thin uid layer only via
conduction. Thus the wall heat ux is estimated by:

kt Prl

kl Prt

lt
ll

The formula presented above, presupposes swirl dominated


ows and is used in the absence of other data. The viscosity ratio
of Eq. (5) is calculated from the formula:




lt
jr n 1  exp 2ajrn
ll

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

and

r n

2.4. Combustion

lw

rn

qdrn

where j = 0.41 is the von Karman constant, a = 0.06 and rn is the


normal distance from the wall. The characteristic velocity is considered to be proportional to the engine speed. The proportionality
constant is taken equal to 0.10 for the purposes of the present study
and was chosen to match the cylinder pressure trace during compression, combustion and expansion.
2.3. Mass transfer
In the WMT mode, in which mass transfer in considered for,
mass ows between zones to maintain the pressure uniform inside
the combustion chamber, which is the actual case in reciprocating
engines. The cylinder pressure is calculated at each CA step using
the ideal gas relation for all the zones:

Pcyl V i

mi
Ru T i ;
MW i

i 1; z

10

Solving (10) for the mass of each zone and summing for all the
zones, we end up to:
z
X

1195

mi mcyl

i1

z
X
Pcyl V i MW i
Ru T i
i1

11

i1

12

Ti

At each CA the mass of each zone is calculated using the equation of state for an ideal gas:

mi Pcyl

V i MW i
;
Ru T i

i 1; z

13

The mass change of each zone during the CA step is therefore:


CA
Dmi mi  mprevious
;
i

i 1; z

14

The mass change of each zone equals the net mass ow from its
neighbouring zones. Taking into account the conguration of the
zones, mass ows only between neighbouring zones:
flow
Dmi mflow
i1!i  mi!i1

15

where Dmi is the total mass variation of zone i and


mass ow from zone i to zone j.
Subject to the boundary conditions:

Dm1 mflow
1!2

mflow
i1!i

is the

since mass does not ow through the wall and

17

The total mass exchange is obtained as follows:

Dmcyl

z
X
i1

Dmi

mflow
1!2

j 1; 84 and i 1; z

19

where Yj is the mass fraction of species j in zone i, xj is the molar rate of production of species j (moles/(cm3 s)), Wj is the molecular weight of species j, q is the density of zone i (g/cm3) and z is
the number of zones.
At each time step, rst the change of composition of each zone
is determined assuming no mass exchange between them. Then
mass is transferred between zones to keep the pressure uniform
throughout the cylinder. Therefore, the change of composition of
each zone is the result of combustion and mass transfer from its
neighbouring zones in the mode where mass transfer is considered
for.

The formation rate of CO and HC in each zone is estimated using


the chemical reactions involved in the Chalmers reduced mechanism for isooctane oxidation [28]. Since it is believed that CO and
HC emissions are largely affected by the existence of the crevice
volumes and the temperature distribution inside the combustion
chamber, the interaction of the crevices with the outer- and therefore colder-zones is considered. To implement this interaction, exchange of species is considered between the zones including the
transfer of combustion products in and out of the crevices.
The formation of nitrogen oxides (NOx), i.e. NO and NO2, is accounted for in the reduced set of chemical reactions created at
Chalmers University [28]. This set includes the chemical reactions
listed in Table 1. The NOx formation mechanism includes the extended Zeldovich mechanism (reactions 13) for NO formation,
the reactions leading to NO2 (reactions 47) and reactions including N2O, which have been found to play an important role in NOx
formation [29].
2.6. Constant zone mass mode description

16

Dmz mflow
z1!z



dY j 
x_ j W j 

;
dt i
q i

2.5. Pollutant formation

The mean cylinder pressure is thus estimated from:

mcyl Ru
Pcyl Pz V MW

Combustion is described using the set of chemical reactions created at Chalmers University for isooctane [28] consisting of 84 species and 412 reactions. These reactions describe the oxidation of
isooctane and the formation of combustion products. Soot emissions are not included due to the premixed nature of combustion.
The rate of production (or destruction) of each species is calculated
and the set of differential equations obtained is solved using the
Chemkin [22] libraries to determine the variation of mixture composition for each zone:

z1

X
flow
flow

mflow
i1!i  mi!i1 mz1!z 0
i2

18
Since blowby is not described in the model, the previous condition must be satised.
The transfer of species and enthalpy is based on the assumption
that the mass owing from a zone to its neighbouring one has the
thermodynamic properties (i.e. temperature and chemical species
composition) of the zone from which it originates.

The two modes of the multi-zone model used are identical as


regards combustion and heat transfer between zones and to the

Table 1
NOx formation reactions.
1
2
3
4
5
6
7
8
9
10
11
12
13

N + NO = N2 + O
N + O2 = NO + O
N + OH = NO + H
NO + HO2 = NO2 + OH
NO2 + O = NO + O2
NO2 + H = NO + OH
NO + O + M = NO2 + M
N + CO2 = NO + CO
N2O + O = NO + NO
N2O + O = N2 + O2
N2O + H = N2 + OH
N2O + M = N2 + O + M
N2O + OH = N2 + HO2

1196

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

cylinder wall. In the WMT mode, mass transfer is determined as


described in Section 2.3. In the CZM mode, which neglects mass
transfer, the mass of each zone is considered constant after a
user-dened CA has been reached, i.e. species and enthalpy transfer are omitted.
From a computational point of view, in the WMT mode, the volume of each zone is determined at the beginning of each CA step
based on the conguration and the (constant) thickness of the
zones. In this mode, mass transfer results from thermodynamic
considerations, to maintain uniform pressure within the combustion chamber (Section 2.3).
In the CZM mode, the mass of each zone remains constant and is
therefore known at the beginning of each CA step. The volume of
each zone is subsequently determined using thermodynamic considerations to maintain uniform pressure. The thickness is thereafter calculated based on the volumes and the conguration of the
zones. In the CZM mode the conguration of the zones is identical
as in the WMT mode, but the thickness of each zone varies.
3. Engine specications operating and modeling conditions
Although the main focus of the paper is to investigate the effect
of mass transfer on emissions formation using a multi-zone model,
a comparison of the predictions of the model to experimental results is provided for a rudimental validation of the model.
The experimental data used in the present work were provided
by Lund Institute of Technology from a relevant experimental
investigation. The details of this investigation have been published
in [5,6]. The engine considered is a Volvo TD100 Diesel engine, the
specications of which are given in Table 2.
The operating conditions used for the purposes of the present
study, i.e. the study of mass transfer and its effect on emissions,
the results for a naturally aspirated case of isooctane fuel are used
as specied in Table 3. The air was heated to 105 C before entering
the cylinder.
Sixteen (16) zones were used for the multi-zone model simulation, including the crevice zone (zone 1). For the WMT case the crevice volume is considered constant throughout the engine cycle
and estimated to be 1.5% of the clearance volume [30]. The thicknesses of the zones at IVC are given in Table 4.
More zones and a ner zone conguration near the wall, compared to the previous studies (10 zones of equal thickness at TDC
in [1618]), was necessary for the more accurate description of
the thermal boundary layer near the combustion chamber wall,
which affects the emissions formed in this region.

Table 2
Engine data.
Compr. ratio
Bore (mm)
Stroke (mm)
Connect. rod (mm)
EVO (deg BBDC)
EVC (deg BTDC)
IVO (deg ATDC)
IVC (deg ABDC)

Zone

Thickness (mm)

26
79
1015
16 (at TDC)

0.047882
0.239409
0.488194
0.488194

Table 5
Description of simulated cases.
Case

Mass transfer inclusion

WMT (base)
CZM 1st HR
CZM 2nd HR
CZM Expans

Entire closed cycle


Up to 1st heat release initiation (20 deg aTDC)
Up to 2nd heat release initiation (10 deg aTDC)
Up to 2nd heat release completion (+10 deg aTDC)

The temperature and composition at IVC are uniform throughout the cylinder as already mentioned and the wall temperature
was estimated around 430 K. Details for the computational procedure can be found in previous papers of the author [1618]. The
computational time for the case of 16 zones is approximately
40 min on an Intel Core Duo 2 GHz Processor, which is relatively
acceptable for such calculations, i.e. including chemical kinetics.
4. Computational procedure
Table 5 presents the four cases, which were simulated and compared. In the rst case (with mass transfer WMT), which is the
base of comparison, mass transfer is allowed between zones
throughout the closed cycle, i.e. during compression combustion
and expansion. In all CZM cases the multi-zone model runs in
WMT mode only up to a certain point in the closed cycle, after
which the mode switches to CZM. In CZM mode mass and species
transfer is prohibited and zone mass is maintained constant for the
remaining cycle. In particular, case CZM 1st HR includes mass
transfer up to the initiation of the 1st heat release, case CZM 2nd
HR includes mass transfer up to the initiation of the 2nd heat release (10 deg aTDC) and in case CZM Expans mass transfer is
neglected after the completion of main combustion (10 deg
aTDC).
It is evident form the aforementioned, that all CZM cases produce identical results with the WMT case, up to the point where
mass transfer is no longer considered for.
5. Results and discussion

17
120.65
140
260
39
10
5
13

Table 3
Operating conditions.
Fuel
Engine speed (rpm)
Lambda
imepnet (bar)
Inlet conditions
Residual fraction (est.) (%)

Table 4
Zone thickness.

Isooctane
1000
2.95
3.8
Naturally aspirated
5

5.1. Comparison to experimental results


In Fig. 2 the experimental pressure trace and heat release rate
are compared to the ones obtained from the multi-zone model
with mass transfer throughout the closed engine cycle (case
WMT). The comparison between the pressure traces and heat release rates is adequate. The heat release rate and the peak pressure
are slightly overestimated. The overestimation of heat release rate
could be attributed partly to the unavoidable smoothing of the
experimental heat release rate and also to the different heat transfer models used during combustion.
Table 6 presents the corresponding exhaust measured and calculated emissions at EVO. All emissions are underpredicted by
about 6% for CO and 11% for HC emissions. NOx emissions are not
a concern in HCCI combustion since they remain at a very low level
for the case studied.

1197

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

100

90

2000

WMT
CZM 1st HRR
CZM 2nd HRR
CZM Expans

Experimental
Calculated
80

80

1600

1200

40

800

20

400

Pressure (bar)

60

HRR (J/deg)

Pressure (bar)

70

60

50

40

30

0
-30

-20

-10

10

20

30

-10

40

-5

15

Species

Measured (mg)

Calculated (mg)

CO
HC
NOx

0.9682
2.536
3.521  103

0.9082
2.244
2.764  103

5.2. Mass transfer description


The multi-zone model results for the in-cylinder mass ow have
been presented in previous studies by the author [1618], therefore only a brief description is provided herein for completeness.
Mass ow is affected by three processes, i.e. compression, expansion and combustion. The upward piston motion during compression induces mass ow from the core zone to the outer zones.
Expansion reverses the ow and induces mass ow from the outer
zones to the inner zones. As soon as combustion commences in a
zone, mass is transferred from this zone to its neighbouring ones.
The combined effect of these phenomena, determines the direction
of the ow between neighbouring zones. The effect of compression,
combustion and expansion on mass transfer between zones is
shown schematically in Fig. 3.
5.3. Effect of mass transfer on pressure traces and temperature proles
The inclusion of mass transfer affects the predicted pressure
trace, as shown in Fig. 4, whereby the four cases simulated are pre-

sented. Comparing cases WMT and CZM 1st and 2nd HR, it is concluded that earlier ignition and a higher peak combustion pressure
results when neglecting mass transfer after the initiation of combustion. The temperature eld is also affected by the inclusion of
mass transfer. Fig. 5 presents the temperature prole prior to
and during combustion for the two extreme cases, i.e. with the
inclusion of mass transfer (WMT) and neglecting mass transfer
after 1st heat release initiation (CZM 1st HR). The different zone
thicknesses for the two cases, which could induce alterations to
the heat transfer estimations, does not seem to be the source of
this discrepancy since the temperature proles remain essentially
identical up to TDC. Therefore this difference must be attributed to
mass transfer during combustion.
To investigate this hypothesis, the heat release, net heat transfer and net enthalpy transfer rates for each zone, are depicted in
Fig. 6 for the WMT case. The net heat transfer rate of each zone
is positive when more heat is gained than lost from the zone. Similarly, the net enthalpy transfer rate is positive, when the enthalpy

2000
Combustion Chamber Wall

Table 6
Measured and calculated emissions.

5o

1800

4o
1600

3o

1400

2o

1200

TDC
-5o

Mass Flow
Crevice Flow
Piston Motion
Combustion

10

Fig. 4. Comparison of pressure traces for the cases examined.

Fig. 2. Experimental and calculated (case WMT) pressure traces and heat release
rates.

CA deg aTDC

CA deg aTDC

1000

-15o aTDC

Temperature (K)

0
-40

800

2
i-1
1

CZM 1st HR
WMT

600

i+1

400
0

Compression

Combustion

Expansion

Fig. 3. Schematic of mass ow during compression, combustion and expansion.

Distance from Wall (mm)


Fig. 5. Temperature proles comparison for the WMT and CZM 1st HR cases.

1198

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

alters both quantitatively and qualitatively the emissions


prediction.
In order to explain the aforementioned trends, the in-cylinder
HC and CO history is discussed in the subsequent sections.

200
150
100
50

Htransfer net (J/deg)

Qloss net (J/deg)

0
40
0
-40
-80
40
0
-40
-80
-2

CA (deg aTDC)
Fig. 6. Zone heat release, net heat transfer and net enthalpy transfer rate, WMT
case.

inow is higher than the outow. The negative peaks in the net enthalpy transfer diagram occurring between 3 and 5 deg aTDC correspond to the inner zones. This net enthalpy transfer rate
observed during main combustion is comparable and for some
zones even greater than net heat transfer rate. As soon as combustion occurs in a zone, mass ows out of this zone to the neighbouring zones, resulting to a net outow (loss) of enthalpy, which
would have remained in the zone if mass transfer was neglected.
This explains the milder temperature prole and the delayed temperature increase of WMT case as compared to the CZM 1st HR
case.
5.4. Effect of mass transfer on emissions formation
In Table 7 the emissions results for all simulated cases are
shown, both in absolute values (g) and relative (%) to the baseline
case (WMT). A signicant deviation is observed between case WMT
and CZM cases. Neglecting mass transfer after the 1st or 2nd heat
release initiation, results to an underestimation of CO and HC by at
least 30% and at most 53%.
Comparing cases CZM 1st HR and CZM Expans, it is observed
that if mass transfer is neglected prior to 1st heat release, HC emissions are signicantly underestimated, whilst if the constant zone
mass assumption is applied at the beginning of expansion, HC
emissions are overestimated. Therefore, the timing at which the
constant zone mass assumption is applied to the multi-zone model

5.4.1. Effect of mass transfer on HC emissions formation


To clarify the effect of mass transfer on HC emissions, a comparison of the in-cylinder HC history for the four cases examined is
presented in Fig. 7. It is observed that when mass transfer is neglected after the 1st or 2nd heat release initiation, the amount of
in-cylinder HC present during expansion is low and essentially
constant throughout the expansion stroke, i.e. HC oxidation does
not seem to take place during expansion.
Neglecting mass transfer after the completion of combustion
(CZM Expans), results to an overestimation of HC relative to the
WMT case, due to the apparent lower HC oxidation rate during
expansion.
The source of these deviations from the baseline case can be located by shifting the focus to the zone HC distribution during the
latter stages of combustion and the early stages of expansion,
which is provided in Fig. 8.
The rst observation is that zones 1, 2 and 3, i.e. the outmost
zones, are the ones mainly responsible for the HC contained in
the cylinder immediately after combustion (10 deg aTDC) in all
cases. The accumulation of mass in zones 1 and 2 during the latter
stages of compression and during combustion observed when
mass transfer is included during these periods (WMT case), is not
captured by the CZM 1st HR and CZM 2nd HR cases. This is the
main reason for the low HC amount contained in zones 1 and 2
in cases CZM 1st HR and CZM 2nd HR. The difference in the HC
amount between these two cases is attributed to the timing of
the constant zone mass assumption; in case CZM 2nd HR more
mass has been accumulated in zones 1 and 2 relative to case
CZM 1st HR.
Another observation is that neglecting mass transfer apparently
affects the HC oxidation rate during expansion. As soon as mass
transfer is prohibited, the HC amount of zones 1 and 2 remains
unaltered, due to the low zone temperature, whilst in zone 3 almost all HC are oxidized. This prohibits further HC oxidation and
produces zero-slope HC history for all CZM cases during expansion
(Fig. 7). It also justies the lower HC amount found in case WMT
relative to case CZM Expans, since the HC transfer from the colder
zones to the hotter ones, and their subsequent partial oxidation is
only captured in the WMT case.

8.0x10-3

6.0x10

HC (g)

HRR (J/deg)

250

Mass Transfer
CZM 1st HR
CZM 2nd HR
CZM Expans

-3

4.0x10-3

2.0x10-3

Table 7
HC and CO emissions for the cases studied.
Case

HC (g)

HC reduction
rel. to base (%)

CO (g)

CO reduction
rel. to base (%)

WMT (base)
CZM 1st HR
CZM 2nd HR
CZM Expans

2.24E03
1.05E03
1.55E03
2.44E03

53
31
+9

9.08E04
5.04E04
4.88E04
6.70E04

45
46
26

0.0x100
5

10

15

20

25

30

35

40

45

CA (deg aTDC)
Fig. 7. Total in-cylinder HC for the cases examined.

50

1199

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

1.0x10-3

0.002

2
1

0.001

4...

WMT

8.0x10-4

2
3

CO (g)

0
0.002
CZM 1st HR

6.0x10-4

4.0x10-4
WMT
CZM 1st HR
CZM 2nd HR
CZM Expans

HC (g)

0.001

2.0x10-4

0.0x100

0.002

10

15

CZM 2st HR
0.001

20

25

30

35

40

45

50

CA (deg aTDC)

Fig. 9. Total in-cylinder CO for the cases examined.

2
3

Comparing Figs. 7 and 9 for cases CZM 1st and 2nd HR it is seen
that although the HC amount during expansion is quite different
between the two cases, a corresponding difference is not observed
for CO.
To investigate further the differences between cases, the CO history in each zone is provided in Fig. 10. In the cases where mass
transfer is neglected after the 1st or 2nd heat release initiation,
the contribution of zone 1 to the total CO present is non-existent

0.002

1
CZM Expans
0.001

2
3
0
5

10

15

20

25

30

35

40

45

50

CA (deg aTDC)
1.6E-008
1.2E-008

2
1

8E-009

4...

WMT

2
4

4E-009
0

7 6 5

1.6E-008

CO (kmol)

Cases WMT and CZM Expans are identical up to 10 deg aTDC.


For the CZM Expans case in Fig. 8, it is observed that HC in zone
3 increase just prior to 10 deg aTDC and then decrease as soon as
mass transfer is neglected. This implies that in zone 3 the temperature at this instant is high enough to oxidize the HC present. On
the contrary, in case WMT the HC contained in zone 3 continue
to increase after 10 deg aTDC, due to mass transfer from zone 2.
This leads to the conclusion that during this period, HC ow from
zone 2 to zone 3 prevails and determines the HC amount in zone 3.
Moreover, since the temperature in zone 3 is high enough, partial
HC oxidation must take place for the WMT case. This can also be
demonstrated by studying the CO emissions formation, as is shown
subsequently.

CO (kmol)

Fig. 8. Zone HC distribution vs. CA for the cases examined.

CZM 1st HR

1.2E-008

8E-009

4E-009

7 6 2

1.6E-008

CO (kmol)

CZM 2st HR
1.2E-008

8E-009

4E-009

0
1.6E-008

CZM Expans

CO (kmol)

5.4.2. Effect of mass transfer on CO emissions formation


Fig. 9 depicts the total in-cylinder CO history at the latter stages
of combustion and during expansion. Since the CO values near
5 deg aTDC were high, the y-axis scale was modied to focus on
the differences between cases during expansion.
At about 8 deg aTDC the total in-cylinder CO is similar for the
four cases. However, in cases CZM 1st and 2nd HR the total CO oxidation rate is higher, as implied by the steeper CO reduction.
Accordingly, the nal CO amount is much lower in these cases relative to the base line case (WMT) by about 45%. The CO reduction
in case CZM Expans, although signicant, is much less (26%),
which implies that including mass transfer between zones during
combustion is essential for the prediction of CO emissions.
In all CZM cases a near-zero CO production rate is obtained during expansion, as implied by the zero-slope CO history. This is the
main source for the differences in the nal in-cylinder CO between
cases WMT and CZM Expans.

1.2E-008

8E-009

4E-009
0

1
7
8

6
10

12

5
14

16

18

20

22

24

26

CA (deg aTDC)
Fig. 10. Zone CO distribution vs. CA for the cases examined.

28

30

1200

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201

3E-009
2E-009

WMT
2

1E-009

0
-1E-009
-2E-009
-3E-009
3E-009
2E-009

CO production rate (kmol/deg)

1E-009

2
1

4...
CZM 1st HR

-1E-009
-2E-009
-3E-009
3E-009
2E-009

CZM 2st HR

1E-009

0
-1E-009
-2E-009
-3E-009
3E-009
2E-009
1E-009

CZM Expans

0
-1E-009
-2E-009
-3E-009
8

10

12

14

16

18

20

22

24

26

28

30

CA (deg aTDC)
Fig. 11. Zone CO production rate vs. CA for the cases examined.

and the contribution of zone 2 very low. However, these zones contain signicant amounts of HC (Fig. 8), which is not oxidized as already mentioned. Thus, the large difference in HC observed
between the two cases during expansion is not followed by a corresponding difference in CO.
In cases WMT and CZM Expans, in which mass transfer is taken
into account during combustion, both zones 1 and 2 contain a signicant amount of CO immediately after main combustion (10 deg
aTDC). The CO present in zone 1 in these cases is attributed to CO
transfer (inow) during combustion, since the temperature in zone
1 is low for CO production via HC partial oxidation.
Since CO is both transferred between zones and formed by partial HC oxidation, the net CO production rate in each zone is provided in Fig. 11. A positive net CO production rate implies that
more CO is produced than consumed. It is obvious from Fig. 11 that
more CO is produced in zone 2 when mass transfer is included during combustion. This can be attributed in part, to the higher HC
present in zone 2 in cases WMT and CZM Expans relative to cases
CZM 1st and 2nd HR.
More important, however, is the positive net CO production rate
during expansion in zone 3 and to a lesser extent in the other
zones , which is absent in all CZM cases. Partial oxidation of HC
transferred to the hotter, inner zones from zone 1 and 2, is the
main source of this CO production.
6. Summary and conclusions
In the present study an investigation on the effects of mass
transfer on the formation of the CO and HC emissions of an HCCI
engine was conducted. A multi-zone model running in two modes

was used for this purpose. In the rst mode the model includes
mass transfer between zones, whilst in the second mode mass
transfer is neglected. The switch between modes occurs at a
user-dened point (CA) in the engine closed cycle.
Four cases were examined, differentiated by the point at which
the switch between modes occurs: (i) including mass transfer
throughout the cycle, (ii) including mass transfer up to the initiation of 1st heat release, (iii) including mass transfer up to the initiation of 2nd heat release and (iv) including mass transfer up to
the completion of 2nd heat release.
The comparison of the different cases revealed that neglecting
mass transfer during combustion inuences the temperature eld
due to the signicant enthalpy transfer occurring during this period. As a result, a steeper pressure trace during combustion and a
higher peak combustion pressure are obtained when neglecting
mass transfer for the case examined.
As regards unburned hydrocarbon emissions, neglecting mass
transfer during combustion results to an under-prediction by at
least 30% and at most 53% relative to the case which includes mass
transfer, for the operating conditions studied. The main source of
this deviation is the failure to capture the mass accumulation towards the relatively cold regions of the combustion chamber during
combustion. Neglecting mass transfer after the completion of main
combustion results to an overestimation by about 9%, since the HC
transfer from the colder regions to the hotter ones and their subsequent oxidation during expansion, is not captured in this case.
CO emissions are underpredicted by about 45% when mass
transfer is neglected after the initiation of the 1st or 2nd heat release. This is attributed to the lower CO amount accumulated in
the colder regions during combustion and to the HC transfer from
the cold to the hot regions and its subsequent partial oxidation,
which is not captured when mass transfer is not considered for.
Neglecting mass transfer after the completion of combustion results to an underestimation of the nal CO amount by about 25%
relative to the case which includes mass transfer, since a signicant
amount of CO is produced during expansion via the aforementioned HC transfer and partial oxidation mechanism.
Acknowledgements
I would like to thank Professor Bengt Johansson, Magnus Christensen and Andreas Vressner from the Lund Institute of Technology
for providing the experimental data used in this paper, for their
willingness and for their useful comments. I would also like to
thank Professor D.T. Hountalas of the National Technical Univ. of
Athens for his help in developing the multi-zone model and Dr.
E.G. Pariotis of the National Technical Univ. of Athens for his helpful
technical comments on the present study. To my sister, Dr. G.P.
Komninou, for her support during the preparation of this study.
References
[1] Alkidas AC. Combustion advancements in gasoline engines. Energy Convers
Manage 2007;48:275161.
[2] Soylu S. Examination of combustion characteristics and phasing strategies of a
natural gas HCCI engine. Energy Convers Manage 2005;46:10119.
[3] Peucheret S, Wyszynski ML, Lehre RS, Golunski S, Xu H. Use of catalytic
reforming to aid natural gas HCCI combustion in engines: experimental and
modeling results of open-loop fuel reforming. Hydrogen Energy
2005;30:158394.
[4] Bhave A, Kraft M, Montorosi L, Mauss F. Sources of CO emissions in an HCCI
engine: a numerical analysis. Combust Flame 2005;144:6347.
[5] Christensen M, Johansson B, Einewall P. Homogeneous charge compression
ignition (HCCI) using isooctane, ethanol and natural gas a comparison with
spark ignition operation. SAE paper no. 972874; 1997.
[6] Christensen M, Johansson B, Amneus P, Mauss F. Supercharged homogeneous
charge compression ignition. SAE paper no. 980787; 1998.
[7] Hultqvist A, Christensen M, Johansson B, Franke A, Richter M, Alden M. A study
of the homogeneous charge compression ignition combustion process by
chemiluminescence imaging. SAE paper no. 1999-01-3680; 1999.

N.P. Komninos / Energy Conversion and Management 50 (2009) 11921201


[8] Hultqvist A, Endgar U, Johansson B, Klingmann J. Reacting boundary layers in a
homogeneous charge compression ignition (HCCI) engine. SAE paper no. 200101-1032; 2001.
[9] Easley WL, Agarwal A, Lavoie GA. Modeling of HCCI combustion and emissions
using detailed chemistry. SAE paper no. 2001-01-1029; 2001.
[10] Jia M, Xie M. Numerical simulation of homogeneous charge compression
ignition combustion using a multi-dimensional model. Proc Inst Mech Eng, J
Automobile Eng 2007;221(Part D):46580.
[11] Ogink R, Golovitchev V. Gasoline HCCI modeling: an engine cycle simulation
code with a multi-zone combustion model. SAE paper no. 2002-01-1745;
2002.
[12] Xu H, Liu M, Gharahbaghi S, Richardson S, Wyszynski M, Megaritis T.
Modelling of HCCI engines: comparison of single-zone, multi-zone and test
data. SAE paper no. 2005-01-2123; 2005.
[13] Aceves SM, Flowers DL, Westbrook CK, Smith JR, Pitz W, Dibble RW, et al. A
multi-zone model for prediction of HCCI combustion and emissions. SAE paper
no. 2000-01-0327; 2000.
[14] Aceves SM, Martinez-Friaz J, Flowers DL, Smith JR, Dibble RW, Wright JF, et al.
A decoupled model of detailed uid mechanics followed by detailed chemical
kinetics for prediction of iso-octane HCCI combustion. SAE paper no. 2001-013612; 2001.
[15] Flowers DL, Aceves SM, Martinez-Frias J, Hessel RP, Dibble RW. Effect of mixing
on hydrocarbon and carbon monoxide emissions prediction for isooctane HCCI
engine combustion using a multi-zone detailed kinetics solver. SAE paper no.
2003-01-1821; 2003.
[16] Komninos NP, Hountalas DT, Kouremenos DA. Development of a new multizone model for the description of physical processes in HCCI engines. SAE
paper no. 2004-01-0562; 2004.
[17] Komninos NP, Hountalas DT, Kouremenos DA. Description of in-cylinder
combustion processes in HCCI engines using a multi-zone model. SAE paper
no. 2005-01-0171; 2005.

1201

[18] Komninos NP, Hountalas DT, Rakopoulos DA. A parametric investigation of


hydrogen HCCI combustion using a multi-zone model approach. Energy
Convers Manage 2007;48:293441.
[19] Jia M, Zhao Xie M, Peng Z. A comparative study of multi-zone combustion
models for HCCI engines. SAE paper no. 2008-01-0064; 2008.
[20] Kongsereeparp P, Checkel MD. Investigating the effects of reformed fuel
blending in a methane- or n-heptane-HCCI engine using a multi-zone model.
SAE paper no. 2007-01-0205; 2007.
[21] Kongsereeparp P, Checkel MD. Novel method of setting initial conditions for
multi-zone HCCI combustion modeling. SAE paper no. 2007-01-0674; 2007.
[22] Mixture composition; 2008. <http://www.reactiondesign.com>.
[23] Annand WJD. Heat transfer in the cylinders of reciprocating internal
combustion engines. Proc Inst Mech Eng 1963;177:97390.
[24] Incropera FP, DeWitt DP. Introduction to heat transfer. 3rd ed. New York: John
Wiley & Sons; 1996.
[25] Yang J, Martin JK. Approximate solution one-dimensional energy equation
for transient, compressible, low Mach number turbulent boundary layer ows.
Trans ASME, J Heat Transfer 1989;111:61924.
[26] Kong S, Ayoub N, Reitz RD. Modeling combustion in compression ignition
homogeneous charge engines. SAE paper no. 920512; 1992.
[27] Reitz RD. Assessment of wall heat transfer models for premixed-charge engine
combustion computations. SAE paper no. 910267; 1991.
[28] Chemical reactions for isooctane; 2008. <http://www.tfd.chalmers.se/~valeri/
MECH.html>.
[29] Amneus P, Mauss F, Kraft M, Vressner A, Johansson B. NOx and N2O formation
in HCCI engines. SAE paper no. 2005-01-0126; 2005.
[30] Heywood JB. Internal combustion engine fundamentals. New York: McGraw
Hill; 1988.

You might also like