Sunteți pe pagina 1din 10

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines

(COMODIA 2001), July 14, 2001, Nagoya

(2-18) A Refined Two-Zone Heat Release Model for Combustion

Analysis in SI Engines
A.E. Catania, *D. Misul, A. Mittica and E. Spessa
Dipartimento di Energetica, Politecnico di Torino
c.so Duca degli Abruzzi, 24 10129 Torino, Italy
Key Words: Combustion, Modeling, Diagnostics

ABSTRACT
The present work gives a further contribution to the heat release analysis of the measured cylinder pressure timehistory, that is the most commonly used combustion diagnostic for determining the actual burning rate in spark-ignition
engines. A critical examination was made of the thermodynamic models applied for such an analysis, paying specific
attention to their calibration techniques, to the simplifications adopted and to the correlations used for evaluating the
bulk gas-wall heat transfer. With respect to these latter, the convective surface-averaged heat flux is usually determined
by a quasi-steady application of Newtons law, regarding the heat transfer coefficient as a function of the instantaneous
flow properties as well as of geometric and operating engine variables. In a previous paper, the authors proposed and
applied a more general complex-variable formulation of Newtons law of convection for modeling the instantaneous
surface-averaged heat flux so as to take the unsteadiness effect of the gas-wall temperature difference into account.
A refined two-zone heat release model for cylinder pressure-data reduction was developed and assessed. As in the
former version of the model, the thermodynamic properties of both reactants and products were evaluated from JANAF
tables with a multiple species equilibrium composition calculation performed for the burned zone. The novelties of the
present model include the following improvements and their combined implementation. In addition to the new unsteady
convection model, a CAD procedure was introduced in order to estimate the burned- and unburned-zone heat-transfer
wall areas for assigned geometric features of the flame front, whereas the burned- and unburned-gas volumes are often
used as weighting factors of their related heat fluxes in the global heat-transfer calculation. Furthermore, the energy
conservation equation was applied to the unburned-gas zone instead of the isentropic relation that is commonly used for
evaluating the temperature of the unburned gas. A refinement was also included for the calibration procedure of the
cumulative mass-fraction burned, or of the fuel-energy released, at the end of the flame propagation process. Usually,
the predicted mass-fraction burned after combustion is completed is matched to the measured combustion efficiency by
adjusting the heat-transfer correlation constants. More specifically, the amount by which the theoretical maximum
mass-fraction burned is taken to be less than unity is set equal the unreleased energy fraction determined from either the
measured HC or the measured HC, CO and H2 in the engine exhaust. This lost energy is evaluated as a fraction of the
total fuel energy available. In the proposed heat-release model, the heat-transfer calibration is made through an overall
energy balance of the whole cylinder charge during the combustion process. The unreleased energy predicted at the end
of the flame propagation is correlated to the combustion efficiency determined from the exhaust gas composition.
The new heat-release model was applied to the analysis of the combustion parameters in a multivalve SI engine
fueled by either gasoline or CNG under a significant sample of operating conditions. The model appeared to be an
accurate tool of combustion diagnostic for SI engines, being also relevant to diesel engine application.
.

Gasoline - n = 2600 rpm - bmep = 440 kPa - MBT timing


RAFR = 0.91

RAFR = 1.00

RAFR = 0.91
5.0

4.0

q [MW/m2]

0.8

Calibration
Authors
[12]

xb

0.6
0.4

RAFR = 1.00

Authors
[12]

3.0
2.0
1.0

0.2
0
328

Calibration
qb
qu
qb
qu

348

368

[deg]

388

408 325

345

365

385

405

[deg]

0.0
328

348

368

[deg]

388

408 325

345

365

385

[deg]

(a)
(b)
(c)
(d)
Influence of the calibration procedure on the mass-fraction burned (a,b) and heat fluxes (c,d) from burned and
unburned gases to the chamber wall, at the indicated operating conditions.
290

405

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

included and the heat transfer correlation constants were


calibrated in order to match the predicted mass fraction
burned, after the combustion process was completed, to
the combustion efficiency determined only from HC in
the exhaust gas. Such a procedure is directly affected by
the assumption of frozen composition for the unburned
gases, which is also weaker at the end of combustion.
The model developed in [9] includes several parameters
that have to be determined. An optimization procedure
was used to estimate the parameters through a predictor
and minimizing a least-square objective function of the
prediction error. A pressure signal from a motored cycle
and the knowledge of zero heat release during this cycle
were applied to tune the parameters. The motored cycle
was taken in the skip-fired engine under investigation.
More sophisticated multizone heat release models
were recently applied for analyzing cylinder pressure
data in SI engines ([11,12]). The mass fraction burned
was calibrated by adjusting the heat transfer constants
so that its asymptotic value after the end of combustion
matched the measured combustion efficiency. This was
determined from HC, CO, and H2 in exhaust gases. As
is shown further on, such a calibration procedure can be
inaccurate, especially for rich mixtures.
Almost all of the abovementioned heat release
models use Woschnis formula ([15]) to calculate areaaveraged heat fluxes to the combustion chamber wall.
The bulk gas-wall convective heat transfer in SI
engine combustion chambers has long been studied,
mostly aimed at developing and assessing heat transfer
models based on Newtons convection law ([15-18]).
For different engines and operating conditions, local
transient heat flux measurements at the cylinder head
([19]) and at the piston crown ([20]) were compared to
the values calculated using the correlations of Woschni
([15]) and Eichelberg ([16]). The comparisons showed a
phase lag and an attenuation of the calculated heat flux
with respect to the experimental heat flux versus crank
angle. Although a dependence of the heat flux on the
measurement location has been observed by many
investigators ([19-26]), whereas surface-averaged heat
transfer correlations are usually used for modeling, the
discrepancy between measured and calculated fluxes in
[19,20] can be attributed to the unsteadiness of the bulk
gas-wall temperature difference ([13]).
In [27] Newtons law of convection was modified
to explicitly include a term taking the time dependence
of the bulk gas temperature into account. A systematic
and rigorous approach employing complex variables to
represent quasi-periodic phenomena was applied to
formulate a more general complex-variable based law of
convection for modeling the unsteady gas-wall heat
transfer in the combustion chamber of a real SI engine
([13]). Other semiempirical nonstationary heat transfer
formulations are also reviewed in [13].
Multidimensional models of engine gas-wall heat
transfer have prohibitive computing times, even if they
rely on simple wall function and k- turbulence models
([28]). Heat transfer models based on thermal boundary

INTRODUCTION
The heat release analysis of a carefully measured
cylinder pressure time-history has long been used as a
diagnostic of the combustion process for determining
the mass fraction of the charge burned in spark-ignition
(SI) engines ([1,5]1). The pressure data reduction to that
end is usually done by a zero-dimensional application of
the thermodynamics laws to the cylinder contents ([3]).
Hence, the heat release type models have the advantage
of simplicity. Although they have no spatial resolution,
they have been a significant contribution because they
incorporate the rate processes occurring in the engine.
Such models are powerful investigation tools as is given
evidence by the current interest they retain as well as by
the continuous enhancement they are being subjected to
([6-13]). In order to examine the combustion parameters
in a newly developed multivalve SI engine running on
either gasoline or CNG, a critical analysis of currently
used heat release models was performed by the authors,
paying specific attention to the simplifications adopted,
to the calibration techniques, and to the global gas-wall
heat transfer correlations. Improvements were proposed
for a more accurate evaluation of combustion quantities.
Therefore, a refined heat release analysis method was
developed, tested and applied for cylinder pressure data
reduction in SI engines.

BACKGROUND AND PRESENT WORK


An extensive review of both heat-release and heattransfer models is made in [13]. It is worth summarizing
its main conclusions, with additional considerations.
The method proposed by Rassweiler and Withrow
in [14] to obtain combustion information from cylinder
pressure data has still recently been used for calculating
the mass fraction burned in SI engines ([5,6,8,10]). It is
based on the assumption that the difference between the
measured cylinder pressure and the pressure predicted
from polytropic compression or expansion scales with
the mass burned. As an advantage, the method is easy to
use, but its disadvantage is that the energy changes are
never explicitly determined from the pressure record.
Furthermore, the polytropic index that should cater to
some extent for the heat-transfer effect is somewhat
empirical and involves rough assumptions. In [5,6] the
polytropic index stems from values calculated by least
squares fits to the log pressure versus log volume data
generally over a 40 deg interval. The actually employed
crank angles depend on the ignition timing, the end of
combustion and on the exhaust valve crank angles. The
compression index is used up to TDC and the expansion
index thereafter. The constraint that the cumulative heat
release should be constant after the end of combustion
is set for model calibration by adjusting the heat transfer
multiplier in [5,6,10], whereas the ion-signal reduction
through a spark plug is used for calibration in [8].
A single-zone heat release analysis procedure was
used in [1,4,9]. In [1,4], the crevice flow effect was also
1

Numbers in square brackets designate references.


291

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

layers were also applied for quasi- or zero-dimensional


engine simulation, with k- turbulence and turbulent
combustion sub models ([29,30]). However, these have
not yielded more accurate results than those of heat
release type models.
The present work develops and applies a further
refined heat-release analysis procedure for pressure data
in SI engine. A two-zone model is employed for the
cylinder contents description. As in the previous version
of the model ([13]), the thermodynamic properties of
both reactants and products are evaluated from JANAF
tables by a multiple species equilibrium composition
calculation performed for the burned zone.
The novelties of the model included the following
improvements and their combined implementation. In
addition to the newly developed unsteady convection
model ([13]), a CAD procedure was introduced in order
to estimate the burned- and unburned-zone heat-transfer
wall areas for assigned geometric features of the flame
front, whereas the burned- and unburned-gas volumes
are usually used as weighting factors of the related heat
fluxes in the heat-transfer calculation. Furthermore, the
energy conservation equation was applied to the zone of
the unburned gas instead of the isentropic relation for
evaluating the unburned gas temperature. A refinement
was also introduced for the calibration procedure of the
cumulative mass-fraction burned at the end of the flame
propagation process. As mentioned, the predicted massfraction burned after combustion is completed is usually
matched to the experimental combustion efficiency, by
adjusting the constants in the heat-transfer correlation.
More specifically, the amount by which the theoretical
maximum mass-fraction burned is taken to be less than
unity is approximately equated to the unreleased energy
fraction determined from either the measured HC or the
measured HC, CO and H2 in the engine exhaust. This
lost energy is evaluated as a fraction of the total fuel
energy available. In the proposed heat release model,
the heat transfer calibration was made by means of an
overall energy balance of the cylinder charge during the
combustion process. The unreleased energy at the end
of flame propagation was correlated to the combustion
efficiency derived from the exhaust gas composition.
The new heat release model was applied to the
analysis of the combustion parameters in a multivalve
SI engine fueled by either gasoline or CNG. It appeared
to be an accurate means of combustion diagnostic for SI
engines, being also relevant to diesel engines.

Fig. 1 Burned- (Vb) and unburned- (Vu)


zone volumes in the engine combustion chamber.
form with reference to the time interval dt = d (for
the symbols refer to Nomenclature).
Figure 1 shows a schematic of the engine with the
combustion chamber split in the burned and unburned
gas zones, which are indicated by shaded volumes. The
following equation is obtained by constraining the sum
of the two zone volumes to be equal to the instantaneous cylinder volume:

dVu + dVb = dV

(1)

The application of the mass conservation equation


to the cylinder charge yields:
dm = d (m f + ma + m r ) = dmu + dmb = 0

(2)

and introducing the mass fraction burned x b = mb m :


mb = xb m,

mu = (1 xb ) m

(3)

As can be easily shown, the energy conservation


equation for the unburned zone can be expressed by:
qu Au

p dVu = mu duu p vu dmu =


= (1 xb ) m duu + RuTu m dxb

(4)

The heat exchange between unburned and burned gases


is neglected in comparison to the gas-wall heat transfer
because of the different temporal scales. Hence, qu Au
expresses the total heat flux from the unburned gas to
the combustion chamber wall. The perfect gas law for
the unburned zone takes the following differential form:

HEAT RELEASE MODEL


Following the conventional approach of two-zone
SI engine models, the heat release analysis is based on
applying the mass and energy conservation principles
for the burned and unburned gas zone, supplemented
with the perfect gas law in each zone, the measured incylinder pressure and information on engine geometric
features and kinematics. The crevice effect was shown
to be very low in [1] and therefore it was neglected. The
basic model equations are formulated in a differential

p dVu = Vu dp + (1 xb ) m d (RuTu ) RuTu m dxb

(5)

so that, eliminating pdVu from eqs. (4) and (5), taking


d (uu + RuTu ) = diu into account, one obtains:
qu Au

+ Vu dp = (1 xb ) m diu

(6)

It is Vu = (1 xb ) m RuTu p , for the perfect gas law, and

292

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

diu = c p dTu , for unburned gas frozen composition

combustion chamber for determining the heat transfer


wall areas Au and Ab. A spherical flame-front model was
used to that end. For comparison, averaged heat-flux
calculations were also made using the respective mass
or volume fractions as weighting factors for unburned
and burned gas heat fluxes so that q = qu (1 x b ) + qb x b ,
as was used in [13], and q = qu Vu V + qb Vb V ([12]).

hypothesis, neglecting the temperature dependence of


ich . Therefore, eq. (6) can also be written as:
u

dp

= (1 xb ) m c p dTu RuTu
(7)
p

It is worth pointing out that the isentropic relation


results from eq. (7) if the heat flux qu is neglected.
The energy conservation for the cylinder content
is expressed by:
d
qA
pdV = d (mb ub + mu uu ) =
(8)

= (ub uu ) m dxb + (1 xb ) m duu + mxb dub


qu Au

UNSTEADY CONVECTION MODEL


According to the procedure described in [13], the
heat flux q for each zone is expressed by:

Introducing the surface-averaged heat flux:


q = qu Au A + qb Ab A

(9)

(qu Au + qb Ab )

+ Vdp = (ib iu ) mdxb +


+ (1 xb ) m diu + m xb dib

V m vu p V m RuTu
=
vb v u
RbTb RuTu

d ( pV m RuTu )
+
RbTb RuTu

( pV m RuTu )

d (RbTb RuTu )
(RbTb RuTu )2

n =1

Tn = An cos n t + Bn sin n t
Twn = Dn cos n t + En sin n t

(11)

T0 and Tw0 are time-averaged temperatures over and

the frequency v is proportional to the engine speed.


Introducing the mean piston speed vp = 2Ln, where L is
the stroke and n the engine angular speed, eq. (14) can
be approximated by:

(12)

D d (T Tw )
q = h T Tw + K

vp
dt

and its differential form:


dxb =

Tw = Tw0 + Twn

n =1

From the expression of the combustion chamber volume


V = mu v u + mb v b = m[v u + x b (v b v u )] one obtains:
xb =

(14)

T = T0 + Tn ;

(10)

taking eqs. (1) and (5) into account, the following form
can be derived for eq. (8):
d

n = 1,2,.., N

in which, the gas and wall temperatures, T and Tw


respectively, are given by their Fourier series over the
considered period of the engine cycle:

the differential form of the burned-zone perfect gas law:


pdVb = Vb dp + m xb d (RbTb ) + RbTb m dxb

1 d Tn Twn
q = h T Tw +

n
dt
n =1

(15)

where, D is the bore diameter and K a dimensionless


quantity to be calibrated. The model equation (15) was
extensively tested and validated in [13] by using spatialaveraged heat fluxes experimentally determined either
in a multivalve pent-roof chamber engine similar to that
of the present investigation ([26]), or in a disk-shaped
chamber engine ([23]), for which experimental pressure
time-histories were available under different operating
conditions. The heat transfer coefficient in eq. (15) was
estimated by using Woschnis correlation ([15,18,25]).
Values of K between 0.1 and 0.3 in eq. (15) were found
to be suitable to fit all the heat flux distributions tested.

(13)

The nonlinear equation system (6) (11) and (13),


conjunctly with eq. (12) and the heat-flux expression
given by eq. (15) in the next Section, was numerically
solved for the differential increment of the three variables, namely, the temperature of the unburned zone Tu,
the temperature of the burned zone Tb and the burned
mass fraction xb, for each crank angle interval d. The
thermodynamic properties of the reactants and products,
in the above equations, as well as the concentration of
the different species, were evaluated at the end of every
crank-angle interval d . The thermodynamic properties
of each species were determined using JANAF tables.
For the burned gas zone, a multiple species equilibrium
composition calculation was carried out similar to that
described in [33], taking dissociation into account. At
each crank angle, the cylinder pressure (data), the fuel
composition (data), the air-fuel ratio (data), the residualgas fraction (calculated according to [34]), and the gas
temperatures (calculated from analysis) were used as
input. For evaluating the surface-averaged heat flux of
eq. (9), a CAD procedure was applied to the real engine

HEAT RELEASE CALIBRATION


The heat transfer constants in the present model
were tuned so that the cumulative mass-fraction burned,
after the end of combustion, meets the following energy
balance of the whole cylinder charge from the start (s)
to the end (e) of the combustion process:

Hp

1+ +

d 1 p
1
qA

Vdp +

m p
m
(16)
+ 1 xb iu iu + xb ib ib

xi H p =

)(

e0

e0

The 0 subscript indicates the enthalpy of the unburned


and burned gas at the temperature of combustion start,

293

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

CNG - bmep = 200 kPa - n = 2000 rpm - RAFR = 1.0 - MBT timing
300
SFF model
xb weighted
Vb weighted

250

200

200

150

150

100

100

50

50

0
323

343

363

qb
qb
qb

4.0

0.6
0.4
0.2

0
403

383

5.0
SFF model
xb weighted
Vb weighted

0.8

Ab [cm2]

Au [cm2]

250

Au
Au
Au

xb

Ab
Ab
Ab

q [MW/m2]

300

qu
qu
qu

SFF model
xb weighted
Vb weighted

363

383

3.0
2.0
1.0

0
323

343

363

[deg]
(a)

383

403

0.0
323

343

[deg]
(b)

403

[deg]
(c)

Fig. 2 - Effect of burned and unburned gas-wall heat-transfer areas (a) on mass-fraction burned (b) and heat flux (c).
with the end-of-combustion composition for both gases.
The second term on the left side of eq. (16) represents
the unreleased energy, per unit charge mass, at the end
of flame propagation. This energy can be estimated on
the basis of measured exhaust emissions, with particular
reference to HC, CO and H2. More specifically, if xiex is

TEST ENGINE AND EQUIPMENT


The two-zone heat release model was applied to
the analysis of pressure data in a production SI engine
running on compressed natural gas (CNG) and gasoline
([31]) with a multipoint sequential port injection.
The engine was a multivalve Lancia 2000 16V
VIS (Variable Induction System). It featured a fast-burn
pent-roof combustion chamber, four valves per cylinder
and one centrally located spark plug (Fig. 1). The main
engine specifications are ([31]): bore 84 mm; stroke 90
mm; compression ratio 10.35; displacement 1995 cm3.
The engine was operated on a test bench instrumented
for typical dynamometer experiments. A multipurpose
exhaust-gas analyzer was used for measuring emission
amounts ([32]). The pressure time-history of the gases
in the cylinder was taken by a water-cooled piezometric
transducer located on the engine head ([31]).

the pollutant fraction measured in the flow stream of the


exhaust gases, the following correlation can be written:

xi H p

1+
ex 1 + + ,
1

xi

ex

H pi

(17)

where ex is a less than unity quantity accounting for


further heat release due to the partial products oxidation
during expansion and exhaust. This quantity depends on
the engine operating conditions. It was semiempirically
estimated with the support of a model similar to that
of [35]. The following expression of the combustion
efficiency can be derived from eq. (17):

c = 1

x H
i

Hp

pi

(1 + + )
,

= 1

x
i

ex H p

iex

Hp

RESULTS AND DISCUSSION


Pressure time histories were acquired at different
engine loads, speeds and air-fuel ratios, in both cases of
engine running on gasoline and CNG, at MBT timing.
Figure 2a plots, for burned and unburned zones,
the gas-wall heat transfer areas (Ab and Au) versus crank
angle, predicted with the three following procedures:
- SFF (Spherical Flame Front) model: the burned zone
is assumed to have a spherical front propagating from

(1 + )

(18)

The combustion efficiency definition given in [12] is:


i xi H p
(19)
c = 1
H p (1 + )
ex

Gasoline - bmep = 790 kPa - n = 4000 rpm - RAFR = 1.0 - MBT timing
1

16.0
qb
qb

0.8

2500

xb
0.4
Unburned gas calculation
Energy equation
Isentropic evolution

0.2

356

376

[deg]

396

416

T [K]

q [MW/m2]

12.0

0.6

0
336

3000

Unburned gas calculation


qu
Energy equation
qu
Isentropic evolution

8.0
4.0
0.0
336

2000
1500

Tb
Tb

Unburned gas calculation


Tu
Energy equation
Tu
Isentropic evolution

1000

356

376

[deg]

396

416

500
336

356

376

396

[deg]

Fig. 3 - Combustion quantities obtained by applying the energy or the isentropic equation to the unburned gas.
294

416

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

Gasoline - n = 2600 rpm - bmep = 440 kPa - MBT timing


RAFR = 0.91

RAFR = 1.00

RAFR = 1.06

1
0.8

Calibration
Authors
[12]

xb

0.6
0.4
0.2
0
328

348

368

388

408

325

345

[deg]

365

385

405

323

343

[deg]

363

383

403

383

403

[deg]

HRR [1/deg]

0.05
0.04

Calibration
Authors
[12]

0.03
0.02
0.01
0
328

348

368

[deg]

388

408

325

345

365

[deg]

295

385

405

323

343

363

[deg]

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

Gasoline - MBT timing


n = 2000
rpm - RAFR = 1.0
n = 2000 rpm - RAFR = 1.0
bmep = 200 kPa
bmep = 440 kPa

bmep = 620 kPa


bmep = 790 kPa

bmepbmep
= 440
kPa - RAFR = 1.0
= 440 kPa - RAFR = 1.0
n = 2000 rpm
n = 2600 rpm

bmep = 790 kPa - n = 2600 rpm

n = 3300 rpm
n = 4000 rpm

1
0.8

xb

0.6
0.4
0.2
0

as well as the corresponding temperatures (Tb, Tu) are


plotted vs. crank angle starting from the spark timing. It
is shown that during the expansion the energy equation
yields lower unburned gas temperatures with respect to
the isentropic evolution. This results into a lower qu but
does not substantially affect the burn rate.
Figure 4 shows the calibration influence on the xb,
HRR, qb, qu, Tb and Tu distributions predicted, at the
indicated conditions, for three different relative air-fuel
ratios, namely, 0.91 in Fig. 4a, 1.00 in Fig. 4b and 1.06
in Fig. 4c. The solid dots correspond to the calibration
procedure proposed by the authors (eq. 16), whereas the
empty squares refer to the procedure applied in [12],
that is, xb = c with c expressed by eq. (19). Significant
differences in the results can be observed for rich
mixtures (Fig. 5a). The differences lessen as leaner
mixtures are approached (Figs. 5b,c). For all RAFRs,
the authors calibration procedure yields asymptotic xb
values that are closer than those of [12] to the values
xb = c with c given by eq. (19) but considering only
HC exhaust emissions, according to [1,4]. However, the
calibration method of [1,4] is subject to the limitations
mentioned before. These also support the accuracy of
the authors calibration procedure.
Figures 5,6 show the influence of the engine load
(Figs. 5a,6a), the engine speed (Figs. 5b,6b) and RAFR
(Figs. 5c,6c) on xb and heat-release rate (HRR), at the
indicated operating conditions, for the engine running

the spark-plug seat; at each crank angle, Ab is evaluated


by applying a 3D CAD procedure which takes the real
geometry of the combustion chamber into account. Au is
the complement of Ab to the total chamber area;
- xb weighted: at each crank angle Ab and Au are taken
to be proportional to the mass fraction burned and to its
complement to unity, respectively ([13]);
- Vb weighted: Ab and Au are proportional to the volume
of the burned and unburned zones, respectively ([12]).
Figure 2b compares the xb distributions obtained
by applying the three abovementioned procedures to
evaluate the gas-wall heat transfer area for burned an
unburned zones. The xb distribution with the SFF model
is normalized through eq. (16). The model with the so
calibrated heat transfer constants was then applied to the
same pressure data using either xb- or Vb-weighted heat
fluxes. The three distributions of xb show similar trends,
though slightly higher asymptotic values of xb were
found with the Vb-weighted procedure.
Figure 2c shows a small influence of the heattransfer-area evaluation procedure on heat fluxes qu and
qb. A similar result was obtained for Vu and Vb. The SFF
model was used to obtain the results that follow.
Figures 3 compare the model outcome when the
energy equation or the isentropic relation was applied to
the unburned gas zone. The predicted xb distributions,
heat fluxes for the burned (qb) and unburned (qu) zones

296

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

CNG - MBT timing


n = 2000
rpm - RAFR = 1.0
n = 2000 rpm - RAFR = 1.0
bmep = 200 kPa
bmep = 440 kPa

bmepbmep
= 440
kPa - RAFR = 1.0
= 440 kPa - RAFR = 1.0
n = 2000 rpm
n = 2600 rpm

bmep = 620 kPa


bmep = 790 kPa

bmep = 790 kPa - n = 2600 rpm


RAFR = 0.91
RAFR = 0.96
RAFR = 1.02

n = 3300 rpm
n = 4000 rpm

RAFR = 1.06
RAFR = 1.11

0.8

xb

0.6
0.4
0.2
0

HRR [1/deg]

0.05
0.04
0.03
0.02
0.01
0
0

20

40

60

80

20

40

60

80

20

40

60

[deg after SA]

[deg after SA]

[deg after SA]

(a)

(b)

(c)

80

Fig. 6 - Effects of engine load (a), speed (b) and RAFR (c) on xb and HRR, for CNG operations.
on gasoline (Fig. 5) and CNG (Fig. 6). The abscissa axis
reports the crank-angle interval starting from the
spark advance (SA) setting. The results in Figs. 5a,6a
show that for both fuels the ignition delay grows when
the engine load is decreased. In fact, at MBT timing and
fixed speed, an increase in engine load corresponds to
higher intake pressure values, which in turn determine
higher cylinder pressures at spark timing. Hence, the
pre-flame reaction rate increases according to Arrhenius
law. The HRR distributions in Figs. 5a,6a show peak
values that are lower for CNG with respect to gasoline.
n = 2000 rpm - RAFR = 1.0

This suggests a reduced flame propagation velocity for


CNG. At bmep = 790 kPa, the difference between HRR
peak values for the two fuels is considerably reduced.
This can be ascribed to the fact that gasoline SA was
not set at MBT, due to the knock onset.
The results of Figs. 5b,6b suggest that, in the low
speed range (2000-2600 rpm), the flame propagation
velocity is higher for gasoline than for CNG. At higher
engine speeds (3300-4000 rpm), the HRR distributions
indicate a virtually equal flame propagation velocity for
both fuels. However, at n = 4000 rpm, SA for gasoline

bmep = 440 kPa - RAFR = 1.0

3.2

[ms]

2.4

1.6

0.8
100

300

500

700

900

bmep [kPa]

297

bmep = 790 kPa - n = 2600 rpm

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

is knock limited. Furthermore, the combustion process


for CNG appears to be significantly influenced by the
engine speed. This can be inferred from the fact that the
amplitude and angular position of the HHR peaks are
affected by the engine speed (Fig. 6b).
The results of Figs. 5c,6c show a stronger RAFR
influence on the flame propagation speed for gasoline
(Fig. 5c) than for CNG (Fig. 6c). In the gasoline-fueled
engine, the combustion crank-angle interval appeared to
increase when RAFR is increased.
Figure 7 shows, for both fuels, the influence of the
engine load (Fig. 7a), speed (Fig. 7b) and air-fuel ratio
(Fig. 7c) on the time interval from the spark discharge
to the 10% mass-fraction burned (flame development
time) and on the time interval between the 10% and the
90% mass-fraction burned (rapid-burning interval). The
flame development time sensibly decreases when either
bmep is raised at constant n or engine speed is raised at
constant bmep for both fuels. Actually, an increase in
either engine speed or engine load determines higher
values of the in-cylinder pressure at spark timing, thus
leading to higher pre-flame reaction rates. The rapidburning interval decreases when either bmep is raised at
constant n or engine speed is raised at constant bmep,
showing similar trends to those reported for the flame
development interval. The engine load and speed effect
on the burning interval is more sensible for CNG than
for gasoline. This is consistent with an expected more
relevant turbulence role on CNG combustion process.
The results in Fig. 7c suggest that for CNG the
flame development time is almost independent of RAFR
although it shows a slight increase for RAFR lower than
0.95. For gasoline, the flame development time shows
an increasing trend when RAFR is increased. The CNG
rapid-burning interval has a slight minimum at RAFR
0.96 and shows a lower dependence on RAFR with
respect to gasoline. As far as gasoline operations are
concerned, the rapid-burning interval increases when
RAFR is raised from 0.9 to 1.1. This trend suggests
that, under gasoline operation, the 10-90% period can
further decrease for richer mixtures. These behaviors
are consistent with the results reported in [31], which
indicate the maximum burning rate at RAFR = 0.96 and
RAFR = 0.82 for CNG and gasoline, respectively.

difference into account, yet maintaining the simplicity


of thermodynamic diagnostic methods. The burned and
unburned-zone heat-transfer wall areas were estimated
by means of a 3D CAD procedure for a spherical flame
front. The energy conservation equation was applied to
evaluate the unburned gas temperature. A more accurate
procedure was proposed for calibrating the cumulative
mass-fraction burned at the end of combustion.
The model was used for the analysis of cylinder
pressure time histories taken in a multivalve pent-roof
SI engine fueled by either gasoline or CNG. Different
engine loads, speeds and air-fuel ratios for MBT timing
conditions were considered. The heat release analysis of
these data gave physically consistent results for both
fuels, supporting the validity of the newly developed
combustion diagnostic procedure.
Acknowledgments
Financial support to this work was provided by
MURST under Cofin98 National Project Diagnostics
and Modeling for the Analysis of Fluid Dynamics and
Combustion in IC Engines, by CNR and also by Fiat
Research Center (CRF).

NOMENCLATURE
A: Surface area. Combustion-chamber area
cp: Specific heat at constant pressure
D: Bore diameter
h: Convective heat transfer coefficient
Hp: Lower heating value at constant pressure
i =iT + ich : Specific enthalpy
m: Mass. Cylinder charge mass
p: Cylinder pressure
q: Heat flux, or heat transfer rate per unit area
R: Gas constant
t : Time
T: Temperature
u = uT + uch : Specific internal energy
vp: Mean piston speed
V: Volume. Combustion chamber volume
x: Mass fraction
: Air-fuel ratio
: Residual gas to fuel ratio
c: Combustion efficiency
: Crank angle
, n: Engine angular speed

CONCLUSION

Subscript
a: Air
b: Burned
ch: Chemical
e: end of combustion
ex: Exhaust
f: Fuel

A refined heat release model has been proposed


and applied to the in-cylinder pressure data reduction
for investigation of SI engine combustion parameters.
The model uses a two-zone description of the cylinder
contents with the thermodynamic properties of reactants
and products evaluated from JANAF tables, performing
a multiple species equilibrium composition calculation
for the burned zone. The model novelties included the
following improvements and their combined implementation. A more general complex-variable formulation of
Newtons convection law was applied for modeling the
instantaneous surface-averaged heat flux so as to take
the unsteadiness effect of the bulk gas-wall temperature

i:
r:
s:
T:
u:
w:

Emission product
Residual gas
Start of combustion
Thermal
Unburned
Wall

REFERENCES
[1] Gatowski, J.A., Balles, E.N., Chun, K.M., Nelson,
F.E., Ekchian, J.A., and Heywood, J.B., Heat Release
Analysis of Engine Pressure Data, SAE Trans., Journal
of Engines, Vol. 93, pp. 961-977, 1984.

298

The Fifth International Symposium on Diagnostics and Modeling of Combustion in Internal Combustion Engines
(COMODIA 2001), July 14, 2001, Nagoya

[2] Amann, C.A., Cylinder-Pressure Measurement


and its Use in Engine Research, SAE Paper No.
852067, 1985.
[3] Foster, D.E., An Overview of Zero-Dimensional
Thermodynamic Models for IC Engine Data Analysis,
SAE Paper No. 852070, 1985.
[4] Chun, K.M., and Heywood, J.B., Estimating
Heat-Release and Mass-of-Mixture Burned from SparkIgnition Engine Pressure Data, Combust. Sci. and
Tech., Vol. 54, pp. 133-143, 1987.
[5] Shayler P.J., Wiseman, M.W., and Ma, T., Improving the Determination of Mass Fraction Burnt,
SAE Paper No. 900351, 1990.
[6] Brunt, M.F.J., and Emtage, A.L., Evaluation of
Burn Rate Routines and Analysis Errors, SAE Paper
No. 970037, 1997.
[7] Oppenheim, A.K., Barton, J.E., Kuhl, A.L., and
Johnson, W.P., Refinement of Heat Release Analysis,
SAE Paper No. 970538, 1997.
[8] Daniels, C.F., The Comparison of Mass Fraction
Burned Obtained from the Cylinder Pressure Signal and
Spark Plug Ion Signal, SAE Paper No. 980140, 1998.
[9] Eriksson, L., Requirements for and a Systematic
Method for Identifying Heat-Release Model Parameters, SAE Paper No. 980626, 1998.
[10] Brunt M.F.J., Rai, H., and Emtage, A.L., The
Calculation of Heat Release Energy from Engine
Cylinder Pressure Data, SAE paper No. 981052, 1998.
[11] Egnell, R., Combustion Diagnostics by Means of
Multizone Heat Release Analysis and NO Calculation,
SAE Paper No. 981424, 1998.
[12] Guezennec, Y.G., and Hamada, W., Two-Zone
Heat Release Analysis of Combustion Data and Calibration of Heat Transfer Correlation in an I.C. Engine,
SAE Paper No. 1999-01-0218, 1999.
[13] Catania, A.E., Misul, D., Mittica, A., and Spessa,
E., Unsteady Convection Model for Heat Release
Analysis of IC Engine Pressure Data, SAE Paper No.
2000-01-1265, 2000.
[14] Rassweiler, G.M. and Withrow, L., Motion
Pictures of Engine Flames Correlated with Pressure
Cards, SAE Trans., Vol. 47, pp. 185-204, 1938.
[15] Woschni, G., A Universally Applicable Equation
for the Instantaneous Heat Transfer Coefficient in the
Internal Combustion Engine, SAE Trans., Vol. 76, pp.
3065-3083, 1967.
[16] Eichelberg, G., Some New Investigations on Old
Combustion Engine problems, Engineering, Vol 148,
pp. 463-547, 1939.
[17] Hohenberg, G.F., Advanced Approaches for Heat
Transfer Calculations, SAE Trans., Vol. 88, pp. 27882806, 1979.
[18] Woschni, G., and Spinder, W., Heat Transfer with
Insulated Combustion Chamber Walls and Its Influence
on the Performance of Diesel Engines, ASME Trans.,
JEGTP, Vol. 110, pp. 482-502, 1988.
[19] Alkidas, A.C., Heat Transfer Characteristics of a
Spark-Ignition Engine, ASME Trans., Journal of Heat
Transfer, Vol. 102, pp.189-193, 1980.

[20] Enomoto Y., and Furuhama, S., A Study of the


Local Heat Transfer Coefficient on the Combustion
Chamber Walls of a Four-Stroke Gasoline Engine,
JSME International Journal, Series II, Vol. 32, No. 1,
pp.107-114, 1989.
[21] Overbye, V.D., Bennethum, J.E., Uyehara, O.A.,
and Myers, P.S., Unsteady Heat Transfer in Engines,
SAE Trans., Vol. 69, pp. 461-494, 1960.
[22] LeFeuvre, T., Myers, P.S., and Uyehara, O.A.,
Experimental Instantaneous Heat Fluxes in a Diesel
Engine and Their Correlation, SAE Paper No. 690464,
1969.
[23] Alkidas, A.C., and Myers, J.P., Transient HeatFlux Measurements in the Combustion Chamber of a
Spark-Ignition Engine, ASME Trans., Journal of Heat
Transfer, Vol. 104, pp. 62-67, 1982.
[24] Alkidas, A.C., Puzinauskas, P.V., and Peterson,
R.C., Combustion and Heat Transfer Studies in a
Spark- Ignited Multivalve Optical Engine, SAE Paper
No. 900353, 1990.
[25] Huber, K., Woschni, G., and Zeilinger, K.,
Investigations on Heat Transfer in Internal Combustion
Engines under Low Load and Motoring Conditions,
Proc., XXIII FISITA Congress, ATA, Torino, 1990.
[26] Alkidas, A.C., and Suh, I.-S., The Effects of
Intake-Flow Configuration on the Heat-Release and
Heat-Transfer Characteristics of a Single-Cylinder
Four-Valve S.I. Engine, SAE Trans., Vol. 100, 1991.
[27] Annand, W.J.D., and Pinfold, D., Heat Transfer in
the Cylinder of a Motored Reciprocating Engine, SAE
Paper No. 800457, 1980.
[28] Gilaber, P., and Pinchon, P., Measurements and
Multidimensional Modeling of Gas-Wall Heat Transfer
in a S.I. Engine, SAE Trans., Journal of Engines,
Vol.97, pp. 839-857, 1988.
[29] Morel, T., Rackmil, C.I., Keribar, R., and Jennings,
M.J., Model for Heat Transfer and Combustion in
Spark Ignited Engines and Its Comparison with
Experiments, SAE Paper No. 880198.
[30] Puzinauskas, P., and Borgnakke, C., Evaluation
and Improvement of an Unsteady Heat Transfer Model
for Spark Ignition Engines, SAE Trans., Journal of
Engines, Vol. 100, pp. 345-360, 1991.
[31] Catania, A.E., Misul, D., Spessa, E., and Martorana, G., Conversion of a Multivalve Gasoline Engine
to run on CNG, SAE Paper No. 2000-01-0673, 2000.
[32] Catania, A.E., dAmbrosio, S., Mittica, A., and
Spessa, E., Experimental Investigation of Fuel Consumption and Exhaust Emissions of a 16V Pent-Roof
Engine Fueled by Gasoline and CNG, SAE Paper No.
2001-01-1191, 2001.
[33] Ferguson, C., Internal Combustion Engines,
John Wiley & Sons, 1986.
[34] Fox, J.W., Cheng, W.K., and Heywood, J.B., A
Model for Predicting Residual Gas Fraction in SparkIgnition Engines, SAE Paper No. 931025, 1993.
[35] Korematsu, J.T.K., Study on Oxidation Processes
of Unburnt HC in Exhaust System of Spark Ignition
Engines, SAE Paper No. 931985, 1985.

299

S-ar putea să vă placă și