Sie sind auf Seite 1von 8

Fuel 89 (2010) 23232330

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Detailed approach for apparent heat release analysis in HCCI engines


Morteza Fathi a,1, R. Khoshbakhti Saray a,*, M. David Checkel b
a
b

Faculty of Mechanical Engineering, Sahand University of Technology, Tabriz, Iran


Mechanical Engineering Department, University of Alberta, Edmonton, AB, Canada T6G 2G8

a r t i c l e

i n f o

Article history:
Received 15 September 2009
Received in revised form 21 April 2010
Accepted 22 April 2010
Available online 5 May 2010
Keywords:
Apparent heat release
HCCI
First law
Specic heat ratio
Temperature correction

a b s t r a c t
Homogeneous charge compression ignition (HCCI) combustion is a spontaneous multi-site auto-ignition
of a lean premixed fuelair mixture, which has high heat release rate, short combustion duration and no
evidence of ame propagation. In HCCI engines, there is no direct control method for the time of autoignition. Auto-ignition timing should be controlled in order to make heat release process take place at
the appropriate time in the engine cycle. Heat release analysis is a diagnostic tool which aids engine
experimenters. It facilitates the endeavors being conducted in obtaining a control method by investigating heat release rate and also cumulative heat release. This study can be divided into two parts. First, traditional rst law heat release model which is widely used in engine combustion analysis was presented
and the applicability of this model in HCCI engines was investigated. Second, a new heat release model
based on the rst law of thermodynamics accompanying with a temperature solver was developed and
assessed. The model was applied in four test conditions with different operating conditions and a variety
of fuel compositions, including i-octane, n-heptane, pure NG, and at last, a dual fueled case of NG and
n-heptane. Results of this work indicate that utilizing the modied rst law heat release model together
with a solver for temperature correction will guarantee obtaining a well-behaved and accurate apparent
heat release trend and magnitude in HCCI combustion engines.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Homogeneous charge compression ignition (HCCI) combustion
mode proposed by Onishi et al. [1] is a reliable method that has
been found to produce low NOx levels, near zero soot and improved
CO2 emissions. It provides equal or greater fuel conversion efciencies compared to that of conventional DI diesel combustion [24].
The advantages of HCCI combustion are commonly associated
with its nature being a spontaneous multi-site combustion of a
very lean and premixed fuelair mixture which has high heat release rate (HRR) and no evidence of ame propagation [4,5]. The
high efciency is due to the ability of operating with high compression ratios, lack of throttling losses at part load, lean combustion,
and close to constant volume ideal Otto cycle heat release.
There are some deciencies intrinsic to HCCI combustion which
should be overcome [4,5]: rst, the lack of any direct control method for combustion timing in contrast to control actions like spark
ignition or fuel injection in SI or CI engines, respectively; second,
producing high levels of HC and CO emissions; third, obtaining
an appropriate fueling rate for achieving high engine loads under
mechanical limitations of engine.

* Corresponding author. Tel./fax: +98 412 3459065.


E-mail addresses: morteza.fathi@yahoo.com (M. Fathi), khoshbakhti@sut.ac.ir
(R. Khoshbakhti Saray), dave.checkel@ualberta.ca (M. David Checkel).
1
Tel.: +98 912 230 5837.
0016-2361/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2010.04.030

Since achieving HCCI combustion is generally difcult, it seems


that heat release analysis is required for a manageable HCCI combustion as it relies on auto-ignition. Identifying early stages of heat
release, timing of the main heat release, and maximum rate of heat
release are necessary for a successful control strategy for HCCI
combustion engine. Due to these facts and other benets of heat
release analysis [613], development of a suitable model for predicting apparent heat release, being able to be applied easily while
eliminating erroneous assumptions, is quite necessary.
To date, numerous researchers have studied heat release analysis in both SI and CI engines. One of the earliest simple attempts
was done by Rasswieler and Withrow [14] in the 1930s. RW
method was a simple but yet efcient method based on correlation
between pressure trace and burnt/unburnt fraction of air/fuel mixture using combustion chamber pictures. The method takes the difference between measured pressure and predicted pressure from
the polytropic compression/expansion to be proportional to the
amount of fuel burned.
Heat release models for both SI and CI engines were presented
by Krieger and Borman [15]. Also, their work included the effect of
dissociation of the product species. Their models as stated by
Gatowski et al. [7] strive for accuracy in representing the thermodynamic properties of the in-cylinder charge and involve substantial computations.
Gatowski et al. [7] presented a model as they claimed being the
rst to give a full description of a heat release model which is

2324

M. Fathi et al. / Fuel 89 (2010) 23232330

Nomenclature
A
cv
height
h
hc
n
N
m
_
m
p
pm
Q
R
T
U
u
up
V
Vd
w
dQgross

area, energy ODE constant


specic heat at constant volume
instantaneous cylinder height
enthalpy, step size
convective heat transfer coefcient
polytropic index
engine speed in rpm
mass
mass ow rate
pressure
motoring pressure
heat
specic gas constant
temperature
internal energy
specic internal energy
mean piston speed
volume
displacement volume
work or characteristic velocity
gross cumulative heat release

Greek symbols
specic heat ratio
h
crank angle

simple in approach and accounts for all major loss mechanisms


while keeping the calculation method simple, as a single zone heat
release analysis procedure. In their study, the thermodynamic
properties of the chamber contents were approximated by specifying that the ratio of specic heats is a linear function of temperature. They argued that accuracy of the calculations based on this
assumption is consistent with the accuracy intrinsic to the use of
the single zone model. This model, being almost the basic model
for single zone heat release analysis, is presented and discussed later in the paper to be compared with the modied rst law model.
Other researchers also developed this kind of modeling, but
their works were to optimize existing models in terms of, for
example, the correlations used. These correlations were namely
correlations for heat transfer, specic heat ratio, etc. For instance,
Chun and Heywood [16] argued that better results could be
achieved if the assumption of linearity with temperature for the ratio of specic heats was studied in more details.
Brunt et al. [17] and Brunt and Platts [8] developed two modied single zone models, one for SI engines and one for CI engines,
respectively, to cater for heat transfer.
Reviewing the literature, few researches can be found being
conducted directly on heat release analysis in HCCI engines. In
2005, Hampson [4] developed a simple engine design tool based
on thermodynamics, allowing engine designers to begin with the
peak pressure limit of engine, desired IMEP, and limitations on
pressure rise rate and thus generate an idealized wire-frame cylinder pressure cycle. He used this tool to identify the ideal HRR for
HCCI combustion. He claimed that for low IMEP, this is a single
mode HRR centered at TDC, and for medium and high loads, a second mode of combustion is required to achieve maximum efciency and retain the benets of HCCI combustion. Thus, this
study developed a design method not a heat release model, and
his work is included here, because of proposing a target heat release for HCCI combustion to shoot for.
In this study, due to the lack of a reliable work on heat release
analysis in HCCI engines, the focus is on investigating the applicability of simplied rst law heat release (SFLHR) model in inter-

Subscripts
c
corrected or convection
f
formation
ht
heat transfer integer
i
position or event
P
products
r
reference
R
reactants
s
sensible
Abbreviations
AHRR
apparent heat release rate
aTDC
after top dead center
bBDC
before bottom dead center
CAD
crank angle degree
EEOC
estimated end of combustion
EVO
exhaust valve open
ESOC
estimated start of combustion
HCCI
homogeneous charge compression ignition
HRR
heat release rate
IVC
inlet valve closure
MFB
mass fraction burned
MFLHR modied rst law heat release
SFLHR
simplied rst law heat release

preting HCCI combustion. Due to the simplifying assumptions,


this model cannot well predict heat release trend in such engines.
Thus, a modied rst law heat release (MFLHR) model which puts
aside erroneous assumptions together with a solver for temperature correction, which calculates the heat released from combustion more accurately, has been developed and will be presented.
2. Experimental setup
Experimental data was obtained from University of Alberta engine research facility. All experiments were performed using a
Waukesha CFR (Cooperative Fuel Research) single cylinder research engine coupled to a DC motoring dynamometer. The basic
engine specications for the current study are shown in Table 1.
Table 2 indicates the CFR engine operating conditions used in
these experiments. The engine was maintained at two constant
speeds of 700 and 800 rpm and ran with an open throttle. The intake system included a 2.4 kW heater with a PID temperature controller to preheat intake air when required. A port liquid fuel
injector was installed a short distance upstream of the intake valve
to ensure proper mixing. The lubricating oil was maintained at a
constant temperature using a heater.
Un-cooled EGR was circulated through an external line to enter
the intake plenum between the air heater and the fuel injectors.
Table 1
Engine specications.
Parameter

Specication

Engine model
Engine type
Combustion chamber
Throttle
Bore
Stroke
Displacement
IVC
EVO

Waukesha CFR
Water cooled, single cylinder
Disk cylinder head, at-top piston
Fully open
82.6 mm
114.3 mm
612 cc
34 CAD, aTDC
40 CAD, bBDC

2325

M. Fathi et al. / Fuel 89 (2010) 23232330


Table 2
CFR engine operating conditions for four considered cases.

Fuel
Compression ratio
Speed (rpm)
Intake temperature
(C)
PRF mass ow rate
(mg/s)
NG mass ow rate
(mg/s)
Air mass ow rate
(g/s)
Pmax location
(CAD, aTDC)

Case 1

Case 2

Case 3

Case 4

i-Octane
13.5:1
700
140

n-Heptane
11.5:1
700
100

NG
17.25:1
800
140

n-Heptane + NG
13.8:1
800
120

78.51

79.78

36.45

82.32

21.76

2.377

2.598

4.46

2.50

8.8

2.2

2.5

3.4

to multi-zone models which can probably be more accurate, due


to the fact that they are numerically more efcient, have less complexity and thus are less time-consuming meanwhile commonly
yielding similar results. Such models do not include spatial
variations.
The accuracy results from a heat release model is dependent on
many factors, among which are, the accuracy of the measured incylinder pressure data, the accuracy of determining TDC location,
and the assumptions adopted in and applied to the model. Investigating the last factor mentioned is the main subject of this study. It
will be shown that in special cases such as HCCI engines having
short combustion duration, some assumptions utilized in previously developed SFLHR model are sources of errors. Therefore, in
this study, two approaches in rst law heat release analyses will
be investigated and compared.
3.1. Simplied rst law heat release model

Case 1

100

The SFLHR model, or basic rst law heat release model as Brunt
et al. [17] named, is the result of applying rst law of thermodynamics to the engine combustion chamber charge. This model
was rst fully developed by Gatowski et al. [7] by considering fuel
injection and all the major loss mechanisms including heat transfer
and crevice ows.
The energy balance for a single zone model for an incremental
crank angle interval is [7]:

Case 2

Pressure [bar]

80

Case 3
Case 4

60

40

dQ gross dU s dW

20

-40

-30

-20

-10

10

20

30

CAD, aTDC
Fig. 1. Average cylinder pressure for 100 consecutive cycles for four cases.

For these steady-state experiments, the EGR rate was manually


controlled with a buttery valve.
A BEI rotary incremental encoder with a resolution of 0.1 CAD
was used to monitor engine rotational speed and coordinate the
pressure trace with respect to crank position. Cylinder pressure
was measured with a Kistler 6043A water cooled pressure transducer in combination with a Kistler 507 charge amplier. The pressure transducer was mounted in the cylinder wall close to the
cylinder head.
Experimental data were acquired using a personal computer
running custom Labview software and using three high sampling
rate NI PCI-MIO-16E1 data acquisition cards.
The pressure trace signal was referenced to the intake pressure
at the time of IVC. Proper digital lter was applied to the pressure
trace signal rejecting the high frequency noise.
Four separate experimental conditions were examined. One
with isooctane as fuel, referred to as case 1, a case with normal
heptane as fuel, which has its peak pressure location fallen before
TDC, referred to as case 2, the other with pure NG as fuel, and at
last a dual fueled case with normal heptane and NG. The operating
conditions for each case are shown in Table 2. The coolant temperature for all cases was set to 100 C. To reduce the effects of cyclic
variations, the pressure traces of all cases are the average of 100
consecutive cycles. These average cylinder pressure traces are
shown in Fig. 1.
3. Heat release model
The simple zero dimensional single zone models for apparent
heat release rate (AHRR) analysis are usually used in preference

hi dmi dQ ht

where dW is the work due to piston motion which equals to pdV and
P
hi dmi is the mass exchanges across the system boundary.
By ignoring all mass exchanges including valve leakages (which
is true for this and many other studies which analyze the system
during closed part of the cycle or nonow period i.e. from IVC to
EVO), fuel injection (which is true, since it was assumed that the
fuel is premixed), blowby, and crevice effects, and taking into account the later-mentioned assumptions, the heat release model
commonly used in the literature is obtained.

dQ gross

c
1
pdV
Vdp dQ ht
c1
c1

Eq. (2), gives the apparent gross heat release on the intervals. By
omitting the last term in the right hand side, Eq. (2) gives the net
apparent heat release, which is not an exact one. There are a number of assumptions in developing Eq. (2):
1. The state of the cylinder contents is dened as average properties of the uniform charge.
2. The change in sensible internal energy is assumed to be a function of mean charge temperature only:

U s muT

By considering the nonow period:

dU s mcv TdT

where cv is the mass average specic heat at constant volume


and is assumed to be a function of temperature only.
3. The mean charge temperature is determined by the ideal gas
law:

pV
mR

4. The specic gas constant is assumed to be constant and this


assumption has its root in that the molecular weights of the
reactants and products are nearly identical [7]. With this
assumption:

2326

M. Fathi et al. / Fuel 89 (2010) 23232330

dT

dpV
mR

The heat transfer coefcient is approximated by:


0:2 0:8 0:73

hc 3:4  height

It is clear from thermodynamics that:

cv
1

R c1

By substituting Eqs. (4), (6), and (7) into Eq. (1), the SFLHR model,
i.e. Eq. (2) will be obtained.

Ph

MFBh PiESOC
EEOC

Dpc;i

iESOC Dpc;i

where MFBh is the mass fraction burned at crank angle h, Dpc is the
corrected pressure rise due to combustion and Dpc ; is calculated by
taking the difference between the ring and motoring pressures
and then referenced to the cylinder volume at TDC:

n

Dpc;i pi  pi1 V i1 =V i V i1 =V r

3.1.2. Specic heat ratio model


In applying traditional SFLHR model, the specic heat ratio is the
most important thermodynamic property, which depends on temperature and composition and to a lesser extent on pressure. There
are a number of functions proposed in the literature for this property. Gatowski et al. [7] proposed a straight line t for cT. Brunt
et al. [17] used a second order correlation between gamma and
temperature. Egnell [19] in his multi-zone model, proposed an
exponential function of temperature for the specic heat ratio.
Other researchers [10,20] also developed models for calculating this
property, but the work of Ceviz and Kaymaz [21] was more general.
They evaluated the specic heat ratio functions and compared them
with their own function derived from the equilibrium combustion
model being dependent on both mean charge temperature and
airfuel ratio. The results of their work showed that taking into account this dependency of specic heat ratio on temperature and
airfuel ratio notably optimized heat release analysis.
Here in this study, the equilibrium combustion model of Olikara
and Borman [22] was fully incorporated with the model to evaluate the specic heat ratio and internal energy, using the thermodynamic data of JANAF table. The equilibrium combustion model was
implemented by considering the products being composed of 11
species.
3.1.3. Heat transfer correlation
One of the most important submodels which must be fed into
the heat release model is the model for evaluating heat lost due
to heat transfer. There are a number of models proposed in the literature for this loss term.
In this study, the well-known expression proposed by Woschni
[23] with modications which made it be appropriate for HCCI engines has been used [24]. In this work, the radiative heat loss has
been ignored. The expression used to calculate the heat ux of convective heat transfer is obtained by:

dQ ht hc  A  T charge  T wall

10

where Tcharge is the mean charge temperature and Twall is the temperature of the wall.

w0:8

11

where height is the effective height of the chamber which is equal


to V/A [24], instead of bore which was used in the original
expression.
The characteristic velocity is obtained as follows:

w C 1 up C 2
3.1.1. Mass fraction burned
Mass fraction burned (MFB) is calculated from the analysis of
measured in-cylinder pressure data. Brunt and Emtage [18] compared the performance of ve MFB models proposed in the literature in accordance with their accuracy in burn angles prediction.
As a result, the model developed by Rasswieler and Withrow
[14] became the preferred model. Also, here in this study, this
MFB model was employed which is obtained by:

p T

V dTr
p  pm
pr V r

12

where C1 = 2.28 and C2 = 0 for compression and 5.4  104 for combustion and expansion.
It is seen that in order to make Woschni model a viable one in
HCCI combustion modeling, some modications such as in temperature exponent (0.73), height, and C2 should be applied.
3.2. Modied rst law heat release model
The MFLHR model is achieved by applying the energy conservation equation to the cylinder contents and not utilizing erroneous
assumptions. By starting with the rst law of thermodynamics for
the cylinder charge during closed valve interval and ignoring mass
leakages, it yields:

dU dW  dQ ht

13

where U is the absolute internal energy of the cylinder contents. By


considering the in-cylinder charge to be a homogeneous mixture of
reactants (denoted by the R index) and products (denoted by P index) which are supposed to be ideal gases (it should be noted here
that Gatowski et al. [7] also used this procedure, but there, the contents were fuel vapor and products and the internal energy was assumed to be a function of temperature only):

m mR mP
dmP dmR

14
15

U mR uR mP uP mR us;R uf ;R mP us;P uf ;P

16

By differentiating Eq. (16), the equation for internal energy change


is obtained:

dU uf ;R  uf ;P dmR us;R  us;P dmR mR dus;R  us;P mdus;P


17
Substituting Eq. (17) and the work term into Eq. (13) yields:

uf ;R  uf ;P dmR us;R  us;P dmR mR dus;R  us;P mdus;P


pdV  dQ ht

18

uf ;R  uf ;P is the difference in internal energy of formation between


the reactants and products. The gross heat release for an incremental crank angle interval, which is the energy released due to combustion, is:

dQ gross uf ;P  uf ;R dmR

19

Substitution of Eq. (19) into Eq. (18) yields the MFLHR model:

dQ gross us;R  us;P dmR mR dus;R  us;P mdus;P pdV dQ ht


20
In this equation for the MFLHR model, mR is the mass of reactants at
any instant and equals to:

mR 1  MFB:m

21

and dmR, is the change in mass of the reactant component due to


combustion and is equal to:

dmR m  dMFB

22

2327

M. Fathi et al. / Fuel 89 (2010) 23232330

Here in this model, it is not only assumed the change in sensible


internal energy being a function of temperature but also a function
of composition, too. The difference between this model and the
SFLHR model developed (Eq. (2)), is that in the SFLHR model it
was rst assumed that the change in the internal energy is a function of temperature only to obtain Eq. (2) to be the governing equation of SFLHR model. Then, for rening the results, it was assumed
that the gamma term is a function of composition and temperature. By contrast, here the dependency of internal energy on composition and temperature was applied into the model itself.
Also another assumption which needs to be considered more
carefully is the ideal gas law used for temperature determination
which is the subject of next section.

1. First, the energy Eq. (20) should be rewritten to produce a function of temperature, species composition and time:

3.2.1. Solver for temperature correction


Using the ideal gas law for temperature calculation, the cumulative gross heat release trend will be at fault, even when utilizing
MFLHR model. Investigating these errors, a solver was developed
for the mean charge temperature.
The internal energy of the homogeneous charge including reactants, i.e. fuel vapor and air, and products which were supposed to
be composed of 11 species, namely, CO2, H2O, CO, N2, O2, H2, OH,
NO, O, H, N, according to the model of Olikara and Borman [22],
was evaluated using the thermodynamic data of JANAF table.
The correction of the temperature was achieved by solving the
initial value problem of ordinary differential equation (ODE) of energy by the aid of fourth-order RungeKutta method. The approach
was:

And the ODE of energy equation, by applying the aforementioned replacement and arrangement, will take the following nal form:

uR  uP
p

dV
hAT  T wall 0
dt

23

By replacing the thermodynamic properties and specic heats of


each species with polynomial functions tted from the JANAF
table,

cv i Rua1  1 a2 T a3 T 2 a4 T 3 a5 T 4
h
i
a
a
a
a
 i Ru a1  1T 2 T 2 3 T 3 4 T 4 5 T 5 a6
u
2
3
4
5

dT
A5 T 5 A4 T 4 A3 T 3 A2 T 2 A1 T A6
dt

24
25

26

where A1 to A6 are instantaneous constants of the equation


which are themselves functions of temperature, pressure and
species mole fractions at any instant, obtained from arrangement of polynomial coefcients of overall 14 species together
with their mole fractions and production rates.
2. This produced equation was used to determine the corrected
temperature by fourth-order RungeKutta method:

b - case 2

a - case 1
90

dmR
d
dus;P
mR us;R  us;P m
dt
dt
dt

130

SFLHR without T correction

110
70

Gross HRR [J / Deg]

Gross HRR [J / Deg]

SFLHR with T correction


MFLHR without T correction
MFLHR with T correction

50

30

10

-10
-40

90
70
50
30
10

-30

-20

-10

10

20

-10
-40

30

-30

-20

-10

CAD, aTDC

CAD, aTDC

c - case 3

d - case 4

10

20

30

108
90

G ross H R R [J / D eg]

Gross HRR [J / Deg]

88

68

48

28

50

30

10

-12

70

-10
-40

-30

-20

-10

10

20

30

-40

-30

-20

-10

CAD, aTDC

CAD, aTDC
Fig. 2. HRR obtained from different models for different cases.

10

20

30

2328

M. Fathi et al. / Fuel 89 (2010) 23232330

dT
fcnt; T and T 0 set point temperature TIVC
dt

bustion durations, results in the correct trend and magnitude in


the heat release diagrams for this case.
In Fig. 2b, the HRR and in Fig. 3b, the corresponding cumulative
gross heat release for case 2 is shown. The two-phase nature of nheptane combustion can be seen in these diagrams. Again, it can be
seen that the correct trend can be obtained, only when the MFLHR
model accompanying with the temperature correction procedure is
used. These gures and also Figs. 2a and 3a show that applying the
wrong model for apparent heat release gives erroneous results not
only in diagrams trend, but also in the low magnitude of peak heat
release rate and cumulative heat release in these cases.
Figs. 2c and 3c show the HRR and cumulative gross heat release
for case 3, respectively. The above mentioned results can be seen in
these two gures, too. But, in this case, the magnitude of peak HRR
and cumulative heat release has been lowered due to applying
temperature correction. Also, it can be seen that applying MFLHR
model together with the temperature correction solver has diminished the negative temperature combustion trend which should
not be seen in the combustion of NG.
The diagrams for HRR and cumulative heat release are shown in
Figs. 2d and 3d, respectively, for the fourth case. Again, it is obvious
that MFLHR model when accompanied with the temperature correction procedure, gives acceptable results, both in terms of shape
and magnitude.
The comparison between mean charge temperatures calculated
from ideal gas law and temperature correction procedure are demonstrated in Fig. 4. It can be seen that the correction procedure for
temperature has smoothen the temperature diagram, lowered its
peak temperature, and increased its post combustion temperature,
for case 1. For cases 2 and 4, this procedure has increased both the
peak and post combustion temperature. Applying temperature correction procedure has totally lowered the temperature for case 3.

27
where the step size equals to:

0:1
6N

28

This procedure was applied to the model from IVC to EVO. It can be
seen from the results, that this corrected temperature gives acceptable results. It is also worth mentioning here, that the CPU time
required for the model to be executed was less than 30 s.

4. Results and discussion


Fig. 2a shows the HRR and Fig. 3a illustrates the corresponding
cumulative gross heat release for case 1. It can be seen from these
gures that applying the traditional SFLHR model to HCCI engines,
which have short combustion durations approximately centered at
TDC, produces large errors. It is well known that the gross HRR
should not be negative after the completion of combustion, and
consequently the cumulative gross heat release should not decrease following the end of combustion. But as can be seen in these
gures, there are negative heat release gradients following the end
of combustion, not only when applying SFLHR model but also in
the case of utilizing MFLHR model without temperature correction,
which is obviously wrong. These gures indicate that there is no
evidence of negative heat release gradient after the completion of
combustion when applying MFLHR model together with the temperature correction solver. Thus, it is guaranteed that applying
the developed model to HCCI engines which have very short com-

b - case 2

a - case 1

Gross cumulative heat release [J]

Gross cumulative heat release [J]

SFLHR without T correction

480
SFLHR with T correction
MFLHR without T correction

380

MFLHR with T correction

280

180

80

-20
-40

-30

-20

-10

10

20

490

390

290

190

90

-10
-40

30

-30

-20

CAD, aTDC

-10

10

20

30

10

20

30

CAD, aTDC
d - case 4

c - case 3
440

490

Gross cumulative heat release [J]

Gross cumulative heat release [J]

490

390

290

190

90

390
340
290
240
190
140
90
40

-10
-40

-30

-20

-10

CAD, aTDC

10

20

30

-10
-40

-30

-20

-10

CAD, aTDC

Fig. 3. Gross cumulative heat release obtained from different models for different cases.

2329

M. Fathi et al. / Fuel 89 (2010) 23232330

b - case 2

a - case 1
1800

Uncorrected Temperature

1700

Corrected Temperature

1500

Temperature [K]

Temperature [K]

1600
1400
1200
1000

1300
1100
900

800

700

600

500

400
-40

-30

-20

1800

-10

10

20

300
-40

30

-30

-20

-10

CAD, aTDC

CAD, aTDC

c - case 1

d - case 4

10

20

30

10

20

30

1600
1600

Temperature [K]

Temperature [K]

1400
1400
1200
1000
800

1200
1000
800
600

600
400
-40

-30

-20

-10

10

20

30

CAD, aTDC

400
-40

-30

-20

-10

CAD, aTDC

Fig. 4. The comparison between temperatures calculated from ideal gas law and temperature correction procedure for different cases.

5. Conclusions
In this study, the applicability and accuracy of traditional SFLHR
model applied to HCCI engines have been investigated. To achieve
this assessment, several specic heat ratio models proposed in the
literature were applied to this model and the results of the most
accurate are shown in the results, and four cases were evaluated.
Results show that this model cannot well and fully predict the
accurate apparent cumulative gross heat release trend and magnitude. To overcome this deciency, the MFLHR model accompanying with a solver for correction of temperature was proposed and
investigated. The results from this model, for all the various working conditions and fuel compositions of evaluated cases, show
good agreement with the concept that states there should not be
any indication of negative gradient in gross heat release.
These results indicate that in HCCI engines, utilizing traditional
SFLHR model results in erroneous results. However, applying
MFLHR model together with temperature correction procedure ensures obtaining more accurate results.
Future work in this paper, the main goal was to present a
model that well predicts the gross heat released from combustion
of in-cylinder contents in HCCI engines. In spite of presenting this
model, research is being conducted by the authors to develop a
more accurate and simpler model taking into account more details
about combustion phenomena in HCCI engines. Also, the authors
are trying to correctly incorporate exhaust gas recirculation in
the model, and investigate its effects on HCCI combustion.
References
[1] Onishi S, Jo SH, Shoda K, Jo PD, Kato S. Active thermo-atmosphere combustion
(ATAC) a new combustion process for internal combustion engines. SAE paper
no. 790501; 1979.

[2] Duret P, Gatellier B, Monteiro L, Miche M, Zima P, Maroteaux D, et al. Progress


in diesel HCCI combustion within the European SPACE LIGHT project. SAE
paper no. 2004-01-1904; 2004.
[3] Sjoberg M, Dec JE. An investigation of the relationship between measured
intake temperature, BDC temperature, and combustion phasing for premixed
and DI HCCI engines. SAE paper no. 2004-01-1900; 2004.
[4] Hampson GJ. Heat release design method for HCCI in diesel engines. SAE paper
no. 2005-01-3728; 2005.
[5] Hosseini V, Checkel MD. Using reformer gas to enhance HCCI combustion of
NG in a CFR engine. SAE paper no. 2006-01-3247; 2006.
[6] Heywood JB. Internal combustion engine fundamentals. New York: McGrawHill; 1988.
[7] Gatowski JA, Balles EN, Chun KM, Nelson FE, Ekchian JA, Heywood JB. Heat
release analysis of engine pressure data. SAE paper no. 841359; 1984.
[8] Brunt MFJ, Platts KC. Calculation of heat release in direct injection diesel
engines. SAE paper no. 1999-01-0187; 1999.
[9] Sastry GVJ, Chandra H. A three-zone heat release model for DI diesel engines.
SAE paper no. 940671; 1994.
[10] Lanzafame R, Messina M. ICE gross heat release strongly inuenced by specic
heat ratio values. Int J Automot Technol 2003;4:12533.
[11] Arregle J, Lopez JJ, Garcia JM, Fenollosa C. Development of a zero-dimensional
diesel combustion model. Part 1: analysis of the quasi-steady diffusion
combustion phase. Appl Therm Eng 2003;23:130117.
[12] Egnell R. The inuence of EGR on heat release rate and NO formation in a DI
diesel engine. SAE paper no. 2000-01-1807; 2000.
[13] Asad U, Zheng M. Fast heat release characterization of a diesel engine. Int J
Therm Sci 2008;47:1688700.
[14] Rasswieler GM, Withrow L. Motion pictures of engine ames correlated with
pressure cards. SAE Trans 1938;42:185204 [Reprinted SAE paper no. 800131;
1980].
[15] Krieger RB, Borman GL. The computation of apparent heat release for internal
combustion engines. ASME paper no. 66-WA/DGP-4; 1966.
[16] Chun KM, Heywood JB. Estimating heat-release and mass-of-mixture burned
from spark-ignition engine pressure data. Combust Sci Technol
1987;54:13343.
[17] Brunt MFJ, Rai H, Emtage AL. The calculation of heat release energy from
engine cylinder pressure data. SAE paper no. 981052; 1998.
[18] Brunt MFJ, Emtage AL. Evaluation of burn rate routines and analysis errors. SAE
paper no. 970037; 1997.
[19] Egnell R. Combustion diagnostics by means of multi-zone heat release analysis
and NO calculation. SAE paper no. 981424; 1998.

2330

M. Fathi et al. / Fuel 89 (2010) 23232330

[20] Klein M, Eriksson L. A specic heat ratio model for single-zone heat release
models. SAE paper no. 2004-01-1464; 2004.
[21] Ceviz MA, Kaymaz I. Temperature and airfuel ratio dependent specic heat
ratio functions for lean burned and unburned mixture. Energy Convers
Manage 2005;46:2387404.
[22] Olikara C, Borman G. A computer program for calculating properties of
equilibrium combustion products with some applications on IC engines. SAE
paper no. 750468; 1975.

[23] Woschni G. A universally applicable equation for the instantaneous heat


transfer coefcient in the internal combustion engine. SAE paper no. 670931;
1967.
[24] Filipi ZS, Chang J, Guralp OA, Assanis DN, Kuo T-W, Najt PM, et al. New heat
transfer correlation for an HCCI engine derived from measurements of
instantaneous surface heat ux. SAE paper no. 2004-01-2996; 2004.

Das könnte Ihnen auch gefallen