You are on page 1of 11

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A.

Tel: (724) 776-4841 Fax: (724) 776-5760


SAE TECHNICAL
PAPER SERIES
2002-01-0073
Development of a Zero-Dimensional Heat
Release Model for Application to Small
Bore Diesel Engines
Peter Schihl, John Tasdemir, Ernest Schwarz and Walter Bryzik
U.S. Army TACOM RD&E Center
SAE 2002 World Congress
Detroit, Michigan
March 4-7, 2002
Quantity reprint rates can be obtained from the Customer Sales and Satisfaction Department.
To request permission to reprint a technical paper or permission to use copyrighted SAE publications in
other works, contact the SAE Publications Group.
ISSN 0148-7191
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely
responsible for the content of the paper. A process is available by which discussions will be printed with the paper if it is published in
SAE Transactions. For permission to publish this paper in full or in part, contact the SAE Publications Group.
Persons wishing to submit papers to be considered for presentation or publication through SAE should send the manuscript or a 300
word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.
Printed in USA
All SAE papers, standards, and selected
books are abstracted and indexed in the
Global Mobility Database
2002-01-0073
Development of a Zero-Dimensional Heat Release Model for
Application to Small Bore Diesel Engines
Peter Schihl, John Tasdemir, Ernest Schwarz and Walter Bryzik
U.S. Army TACOM RD&E Center

ABSTRACT
A zero-dimensional heat release model has been
formulated for small bore, automotive-type, direct
injection diesel engines and compared with high-speed
data acquired from a prototype single-cylinder engine.
This comparison included a significant portion of the full-
load torque curve and various light-loads with variable
speed, injection timing sweeps, and injection pressures.
In general, the agreement between the predicted net
heat release rate profiles and the experimentally,
indirectly-determined profiles was acceptable from a
mean cylinder pressure point-of-view while employing a
single constant for the turbulent mixing dissipation rate.
The proposed model also revealed that moderate swirl
rates included in this study had little impact on the gross
fuel burning rate profile especially at higher load
conditions.
INTRODUCTION
The heat release event in direct-injection diesels is a
complex, non-linear phenomenon that usually is
modeled on a dimensional basis according to the
application. Such analysis includes two- and three-
dimensional, chemically reacting CFD (computation fluid
dynamics) approaches such as KIVA[1-3], quasi-
dimensional models that divide the combustion chamber
into a number of reacting and non-reacting zones[4-17],
and empirical, zero-dimensional models[18-22].
Standard thermodynamic, engine cycle simulation codes
include the latter two types of models since this
particular application is typically employed for predicting
global performance quantities even though it is possible
to incorporate joint CFD-cycle simulation design
approaches at the cost of additional required
computational time.
Standard zero-dimensional models employed within
such cycle simulation codes are based on empirical
burning laws that include a number of engine-dependent
constants whose origin are related to the injection
system, speed operation, combustion chamber
geometry, and coolant system[18]. A database of engine
specific constants maybe utilized to predict the
performance of future diesel engines if the heat release
model predictions are within reasonable tolerance of the
actual combustion process. Translation from such a
database requires careful consideration of variances
from operating speed, compression ratio, piston bowl
shape, and injector actuation, control, and nozzle
geometry. This approach is limited in application for
predicting emissions due to the dimensionless spatial
environment.
Multi-zone models typically are Arrhenius in nature from
a combustion perspective and thus require one or more
engine-dependent constants. Each zone is nearly
independent since heat transfer is modeled as a global
phenomenon that is redistributed locally based on zone
energy content and thus local heat diffusion is not
addressed in a manner that simulates rapid flame
spread during the premixed burn[7]. The inclusion of
additional zones in comparison to zero-dimensional
models does allow for a more detailed investigation of
the interaction between the spray and combustion
chamber that is nearly impossible with zero-dimensional
models and thus provides an opportunity to also predict
emissions.
Multi-dimensional CFD codes commonly utilize a
characteristic time scale combustion model comprised of
some combination of local chemical kinetic and mixing
times[3,23]. Each scale is determined based on local
turbulent mixing and reactivity that are in turn based on
the chosen turbulence and spray formation submodels.
Again, combustion constants are tuned for particular
engine operating ranges due to the sensitivity of break-
up and turbulence submodels on engine speed and
load. CFD approaches require more computational time
in comparison to multi-zone or zero-dimensional models
but are most relevant for studying detailed physics
associated with in-cylinder combustion behavior and
thus are most applicable toward the study of emission
formation mechanisms.
The aim of the proposed model is to bridge the gap
between such empirical constants and the associated
physics of the heat release event with the ultimate goal
of providing reasonable a priori prediction of engine
performance prior to prototype build at the cost of
minimal empiricism. This proposed model is an
extension of an earlier model developed for quiescent,
large bore diesels[22], but with the addition of a swirl
submodel for combustion heat release rate impact and a
K- turbulence submodel for wall heat transfer
calculations.
EXPERIMENTAL ENGINE
The engine utilized in this study was a single cylinder
(SCE) automotive-type diesel[24] as shown in table 1.
Particular features of this engine are EGR (exhaust gas
recirculation) capability, four-valve variable swirl head
with short and long runners, and a high pressure,
piezoelectric controlled common rail injection system.
Instrumentation included in-cylinder pressure, injector
needle lift, standard low speed measurements, and
various gas analyzers. The test plan was predominately
focused on light-load operating conditions with
numerous perturbations of injection timing, rail pressure,
EGR rate, and pilot timing and dwell. Only a portion of
this data will be reported in this paper and all cases
include zero EGR.
Table 1 : CIDI
1
Engine Specifications
Engine Parameter Description
Injection system FEV CORA II
Rail pressure (bar) 500 1200
Nozzle geometry (mm) 6 x 0.124
Bore x stroke (mm) 70 x 78
Connecting rod (mm) 149
Compression ratio 19.5
Swirl number 2.4
2

Displacement (cc) 300
Operating speeds (rpm) 1500 3000
IMEP range (bar) 3 18
Boost system Shop air
1
Compression Ignition Direct Injection

2
Flow bench demonstration at maximum valve lift


Heat release analysis was performed using standard
thermodynamic first law analysis and the ideal gas law.
The specific heat ratio was calculated based on an ideal
gas mixture of CO
2
, H
2
O, N
2
, O
2
and gaseous diesel fuel
when appropriate, i.e. after start of injection, and the
bulk cylinder temperature was determined based on
corrected real gas behavior[25] and estimated in-
cylinder charge mass the corresponding equation of
state is given below:
3
17 . 3
26
2
92 . 2
18
49 . 2
10
64 . 1
3
3
) 100 / (
10 95689 . 2
) 100 / (
10 52579 . 6
) 100 / (
10 34248 . 4
) 100 / (
10 50053 . 8
10 09059 . 1
P
T
x
P
T
x
P
T
x
T
x
x
P
RT

+
+ + =
(1)

Each specie mole fraction was initialized at a chosen
time following intake valve closure and thus, the heat
release calculations can account for residual gas effects.
Furthermore, the calculated heat release rate at a given
crank angle was employed in conjunction with a single
step, global C
n
H
m
chemistry model to determine
perturbations in the specie mole fractions upon initiation
of the injection process. Since the apparent heat release
rate does not differentiate between heat transfer and
gross burning rate, and typical combustion efficiencies in
diesel engines are 99%, a speed up factor was added to
the chemistry model to ensure a nearly complete burn
and thus a more accurate calculation of the charge
specific heat ratio. The pressure and volume derivatives
were calculated based on a two point, forward
differences scheme in order to minimize any potential
damping of the higher frequency combustion (premix
burn) phenomenon that might be difficult to capture due
because of encoder resolution (0.5 crank angle).
Additionally, all experimental pressure traces were
conditioned with a digital low pass filter that had a cutoff
frequency of twice the engine speed preceding heat
release analysis.

A sensitivity study was performed to elicit the influence
of n and m on the global chemistry model in order to
determine if an ideal fuel might be employed as a
surrogate for a heavier hydrocarbon fuel. This study
revealed that the heat release rate profile is insensitive
to C/H ratio fuels that might act as surrogates (12 < n <
16) for diesel fuel because of the negligible impact of
fuel type on the specie mole fraction distribution. N
2

tends to dominate the specific heat ratio calculation and
small perturbations in CO
2
, H
2
O, or O
2
mole fractions are
of minor consequence on the heat release rate profile.

Table 2 : Test Fuel Properties
Fuel Property Description
Type DF-2
Cetane number 53
C/H mass ratio 6.54
Density (kg/m
3
) 842
Sulfur level (ppm) 400
Molecular weight (g/mol) 212
Heating value (MJ/kg) 42.8

MODEL DESCRIPTION
The approach for modeling the heat release event
focuses on capturing the large-scale mixing behavior
and then relating this energy to characteristic, two-stage
combustion submodels. This bulk motion is determined
through the superposition of the transient solutions for
bulk tumble and swirl within one spray sector. The
energy associated with each large-scale motion includes
contributions from the injection process, squish, head
flow design, and engine speed. A description of each
submodel is given in the ensuing sections.
The thermodynamic model is single-zone and given by
the following governing conservation of energy and
mass equations:
V
m
m
dt
dm
Q
dt
dm
LHV V P Q
dt
dE
in
evap
b
= =
+ + =


(2)
where is the wall heat transfer, LHV is the fuel lower
heating value, dm

Q
b
/dt is the total gross fuel burning rate,
is the mass rate crossing the boundary, V is the
combustion chamber volume, is the evaporation
rate, and m is the charge mass. The system energy, E,
is determined based on a mixture of ideal gases
comprised of CO
in
m

evap
Q

2
, H
2
O, N
2
, O
2
, and fuel[26-27]. The
combustion chamber volume, V, is given for a crank-
slider mechanism[28] while the evaporation rate is
assumed to proceed at the injection rate. After initializing
the in-cylinder chemical composition, a global single-
step reaction is employed to monitor production and
destruction of relevant species using pentadecyne
(C
15
H
28
) as the fuel surrogate. The premixed burn is
assumed to occur at a local rich fuel-air ratio that will be
discussed in a subsequent section while the diffusion
burn is assumed to occur at stoichiometry. The resulting
mixture of ideal gases is used to determine the charge
specific heats and thus the temperature derivative.
Afterward, the equation of state given by EQ (1) is
utilized to calculate the bulk in-cylinder pressure.
WALL HEAT TRANSFER A zero-dimensional
turbulence model is utilized to predict the bulk Reynolds
number in order to predict the global heat transfer
coefficient based on analogous turbulent pipe flow. This
approach has been included in past efforts[29-31] as
given by:
2
0
3
2
3307 . 0
3
2 1
5 . 0 5 . 1
B
l
A
V
l
dt
d k
RD
m
k
l
K
C P
m
k
l
RD m P
dt
dk
P
dt
dK
k
b
k
k
b
k
< =
= |
.
|

\
|
= |
.
|

\
|
=
+ = =


where K is the bulk flow energy, A
b
is the cylinder cross-
sectional area, k is the turbulent kinetic energy, P is the
turbulent energy production term, m is the charge mass,
RD is the rapid distortion term, is the dissipation rate,
B is the bore, and C
b
is a constant. A characteristic
turbulent velocity was chosen to determine the global
heat transfer coefficient as shown below:
5 . 0 5 . 0
5 . 0
2
2 2
3
2 2
2
Re Re
|
.
|

\
|
= |
.
|

\
|
=
(
(

|
|
.
|

\
|
+ + =
= =
m
k
u
m
K
U
S
u U U
l U
a Nu
p
ch
k ch b


(4)
where a and b are constants, and S
p
is the mean piston
speed. Radiation is accounted for based on the bulk gas
temperature[32]

i
i rad
T T b Q ) (
4 4


(5)
where T
i
is the wall temperature and b is the bulk gas
wall emissivity.
SWIRL - The swirl rate is determined based on rigid
body bulk flow and is calculated based on assumed
cylinder wall friction losses[28]:
(
(

|
.
|

\
|
+
(
(

|
.
|

\
|
+
=
=
2
4
2
8
1
) (
B
d
h
z
B
d
h
z
B m I
dt
I d
b
b
b
b
c
liner bowl head
s c


(6)
where
s
is the swirl rate, I
c
is the charge mass moment
of inertia, z is the piston position relative to the fire deck,
d
b
is the bowl average diameter, h
b
is the average bowl
depth, and
head
,
bowl
,
liner
are shear stresses[28]. The
initial swirl rate was given according to extensive flow
bench testing[24].
PREMIXED COMBUSTION Upon ignition, the fuel
trapped in the jet shear layer is determined based on the
spray model that is discussed in a later section. A flame
is assumed to propagate from the spray tip toward the
injector tip entraining fresh fuel-air packets into its front
and subsequently consuming this mixture according to
local turbulence and shear layer thermodynamic state.
The flame spread is described according to the
entrainment model of Blizzard and Keck[33] and
Tabaczynski et al.[34] but with modification of the
entrainment rate for an injected diesel jet.
(3)

pb en pb
jet l f
en
m m
dt
dm
U S u A FA
dt
dm

=
+ + = )
~
(
(7)
where m
en
is the entrained fuel mass, FA is the local
fuel-air ratio (linear distribution about mean shear layer
fuel-air ratio), A
f
is the local flame front surface area, U
jet

is the representative local jet velocity, S
l
is the laminar
flame speed, m
pb
is the premixed phase burned fuel, and
is a characteristic Taylor time scale. Each quantity is
defined below:
| |


l u
l S
R R A l u
l l
t
l
t
pm f
~
Re Re
' ) ' (
~
5 . 0
2 2
= =
+ = =


where is the bulk mixing rate, R is the radial distance
from the spray centerline to the onset of the shear layer,
pm
is the premixed fuel-air shear layer thickness (
pm
~
spray tip penetration[35]),
t
is the Taylor scale, and l is
the representative mixing length scale.
MIXING RATE Both the premixed and diffusion
controlled combustion phases are reliant on both the
bulk mixing and effective mixing rates. The bulk mixing
rate is representative of bulk mixing motion and is
modeled accordingly in similar fashion to the eddy
break-up model[36]. Essentially, the bulk mixing motion
is lumped into a representative eddy for each spray
plumeair entrainment event by conserving both mass
and angular momentum. These resulting simplifications
maybe represented as follows:
2 2
tan
2
4
) exp(
) (
8
1
2 2
2
2
3
2
b
s
b
sw
inj
inj
inj s
sq sq
sq s sq inj
e
e
s sw s inj
L S L S
P
S U m
P t a a D
l U m ABS P m m
dt
dm
l m D P P P
dt
d

)
`

|
.
|

\
|
|
.
|

\
| +
=
= =
= + =
= + + =


where is the injection production term, is the squish
production term,

is the dissipation term, m


inj
P

s
P

s D
e
is the
eddy mass,

is the swirl production term,

is the
injection rate, is the squish mass flow rate, l
sw P
sq
m

inj
m
sq
is the
squish length scale, a is the dissipation constant, U
inj
is
the injection velocity, t is the calculation time step, L
b
is
the break-up length, is the spray cone angle, and S is
the spray tip penetration. The eddy mass is assumed to
consist of the charge mass included within the piston
bowl and the projected volume above the bowl while the
squish mass flow rate is determined based on piston
speed, squish area, and bulk density[28] and the squish
length is defined as (z/2 + l/4). Swirl production is only
valid upon establishment of the liquid length core and is
otherwise zero. The representative eddy length scale is
chosen as the following inverse relationship
d z B l +
+ =
1 1 1

where d is a representative bowl depth.
Spray Formation Process The mixing model is
dependent on the injection process due to its momentum
contribution. Choice of a zero-dimensional model
includes numerous options including those proposed by
Wakuri[37], Dent[38], Chikahisa and Murayama[39],
Schihl et al.[40], and Naber and Siebers[41]. In order to
include both near injector and far-field entrainment
effects, the Hiroyasu and Arai model[42] was chosen
and is given below for completeness
(8)
5 . 0
25 . 0
5 . 0
1
) ( 95 . 2 ) (
) 2 (
39 . 0 ) (
t d
P
t S t t
P c
d
t
c
t U
t S t t
o b
o l
b
d
inj
b
|
|
.
|

\
|
= >

= = <



(11)
where t
b
is the break-up time, d
o
is the orifice diameter,
and c
1
are break-up constants, P is the orifice pressure
drop,
l
is the injected fuel density, and c
d
is the
discharge coefficient.
DIFFUSION CONTROLLED COMBUSTION Injected
fuel packets non-inclusive within the premixed shear
layer are allowed to mix according to the large mixing
rate, . Each parcels mixedness is monitored
according to the normalized mixture variable given
below:
(9)

+
=
t t
t
eff
i
i
dt 1
(12)
where
eff
is the effective mixing rate. A parcel is
considered mixed and available for combustion once its
mixedness reaches unity. The effective mixing rate is
defined by the conjugation of bulk motion, wall
impingement, and air utilization and is defined according
to the following characteristic time scales
(13)
util
mix
wall
mix
eff

=
Each time scale ratio will be defined in subsequent
sections. The overall effective mixing rate is utilized to
determine the diffusion phase burn rate and is given by:
eff db a
db
m m
dt
dm
) (

=
(14)
where m
db
is the mass of fuel burned during the diffusion
phase of combustion and m
a
is the available mass of
fuel, i.e. fuel parcels with mixedness of unity.
Wall Effect - Upon impingement of the diesel spray onto
the combustion chamber wall, the local burning rate
tends to slow down due to quench and poor mixing.
Given the proposed model is global in nature, it is
difficult to locally assess the contribution of this
phenomenon on the bulk fuel burning rate. Instead, a
wall - jet film spreading model is employed to estimate
the impact of fuel mass near the wall on the overall
(10)
mixing and burning rate. This characteristic time
approach was formulated based on observations made
by comparison of this proposed submodel with
experimentally determined heat release rate profiles
from two large-bore diesel engines[22]:
900
4500 6500 8500 10500 12500
Pressure at Start-of-Injection (kPa)
T
e
920
940
960
980
1000
1020
1040
1060
m
p
e
r
a
t
u
r
e

a
t

S
t
a
r
t
-
o
f
-
I
n
j
e
c
t
i
o
n

(
K
)
0
1
1

=
+
+
=
imp
imp imp imp
imp imp
imp
wall
mix
L
t U L
L S
L


constant laminar
flame speed
(15)
and t
imp
is the time elapsed after impingement, S
imp
is
the impingement length, and L
imp
is the tangential
spreading length of the jet on the wall that is limited to
the arc length between adjacent impinging fuel jets.
Figure 2 : Influence of Start of Injection
Thermodynamic Condition on Laminar Flame Speed
-20
-10
0
10
20
30
40
50
-10 0 10 20 30
Engine Positon (degree)
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
0
data
N = 1
N = 1.7
N = 2.4
penetration
Air Utilization Effect The availability of oxidizer does
decrease throughout the injection event and can play a
major role in the overall global burning process with
increasing load. Though oxidizer burnout is a local
phenomenon, it is represented in global fashion for the
proposed model according to the following functional
relationship:
2
2
lim
,
1
(
(

|
|
.
|

\
|
=
+
m
m
d p a
util
mix


(16)
where m
a,p+d
is the total fuel mass available for
combustion (including premixed and diffusion phases)
and m
lim
is the maximum fuel mass that may burn at
stoichiometry given the associated global air-fuel ratio.
Figure 3 : Influence of Swirl on the Heat Release
Rate Profile at Full Power.
EXPERIMENTAL RESULTS
The particular speed-load points included within this
study are shown graphically in figure 1. This data set
encompasses twenty different operating points including
timing sweeps at both 2000 rpm and 2500 rpm light load
conditions, and a portion of the full-load torque curve.
Most importantly, all simulation results presented within
this paper are based on utilizing the experimental
ignition delay (defined by the zero cross-over point of the
heat release rate profile) as an input to the combustion
model. This approach was taken to eliminate any
potential biasing associated with predicting inaccurate
ignition delay periods. Additionally, unless noted
otherwise, all simulations included a constant dissipation
coefficient (a) regardless of speed and load changes.
FLAME SPREAD IMPACT One key parameter in fitting
the premixed phase burn rate profile is the laminar flame
speed and thus attention must be paid in judiciously
scaling this parameter as a function of a particular
operating condition. The temperature and pressure at
SOI (start-of-injection), and the mean shear layer
equivalence ratio dictate the mean laminar flame speed.
Due to the lack of available published data for diesel-like
fuels, a methodology was adopted to account for this
key premixed phase combustion parameter. As shown in
figure 1, the temperature and pressure at SOI is similar
at part-load and thus arguably warrants the employment
of a constant laminar flame speed (assuming the mean
shear layer air-fuel ratio is also similar) as a first order
approximation. The full-load cases tend to have much
higher temperatures at SOI (up to 10%) and thus the
flame speed must be modified to account for this
operating condition difference. Additionally, if the laminar
flame speed for diesel fuel is predominately influenced
by temperature in a fashion similar to iso-octane[28]
given comparable mean shear layer equivalence ratios,
then it is appropriate to utilize scaling to determine this
key premixed phase combustion parameter at elevated
0
5
10
15
20
1000 1500 2000 2500 3000 3500
Engine Speed (RPM)
I
M
E
P

(
b
a
r
)
full-load part-load
Figure 1 : CIDI Operating Matrix. IMEP Indicated
Mean Effective Pressure
temperatures. Thus, S
l
/S
l,ref
~ (T/T
ref
)
2

is employed to
calculate laminar flame speeds over the operating
regime included in this study where T
ref
and S
l,ref
are
reference temperature and laminar flame speed.
Pressure can have a significant impact on flame speed
over large variances in operating condition but is less
significant for this limited study.
IMPACT OF SWIRL The influence of swirl on the heat
release rate profile is a function of the injection velocity,
liquid length, charge density, spray angle, and engine
speed. Given that the swirl number was held constant
over this particular operating map it is impossible to fully
assess the capability of the proposed submodel to
match experimental data. Instead, one full load and two
light load conditions were examined to qualitatively
examine the impact of swirl rate on the heat release rate
profile in order to properly choose the effective initial
swirl rate at IVC (intake valve close). For this brief study,
the swirl number was varied from nearly quiescent (n =
1) to that demonstrated at full valve lift on a flow bench
(n = 2.4).
At full power (3000 rpm), swirl rate had a minor impact
on the heat release rate profile. As shown in figure 3,
increasingly higher swirl numbers demonstrated a minor
increase in the mixing controlled phase burn after wall
impingement but was of little significance from a mean
in-cylinder pressure point-of-view. This behavior is
expected given the high injection velocity associated
with this operating condition and the short impingement
length of the CIDI combustion system, i.e. the radial
swirl-induced momentum flux is small compared to the
axial air entrainment rate attributed to the injected
momentum. For case 2, the engine speed was reduced
500 rpm and the load was reduced to a moderate level
such that wall impingement occurred sometime near end
of injection in comparison to the prior rated power case.
As shown in figure 4, increasing swirl rate was more
pronounced during the first half of the mixing controlled
phase, but again had a minor impact on peak cylinder
pressure (< 3%) from a quiescent to the nominal swirl
rate (2.4). Last, the engine speed was cut in half and the
load was chosen at an idle-like condition such that wall
impingement would occur after the end of injection due
to a lower average injection velocity. As shown in figure
5, the varying swirl rate had almost no impact on the
heat release rate profile. This observation maybe
-10
0
10
20
30
40
50
60
70
-10 -5 0 5 10 15 20
Engine Position (degree)
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
-4
+2
-10
-20
-10
0
10
20
30
40
50
60
-10 0 10 20 30
Engine Positon (degree)
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
0
200
data
N = 1
N = 1.7
N = 2.4
penetration
-10
0
10
20
30
40
50
60
70
-10 -5 0 5 10 15 20
Engine Position (degree)
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
-10 -4 +2
Figure 4 : Influence of Swirl Rate on the Heat
Release Rate Profile at 2500 RPM and 5 bar IMEP.
-20
0
20
40
60
80
-10 0 10 20 30
Engine Positon (degree)
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
0
200
data
N = 1
N = 1.7
N = 2.4
penetration
Figure 6 : Influence of Injection Timing Sweep on
Heat Release Rate Profile. Top Figure is for 2000
rpm 3 bar IMEP and the bottom figure is for
2500 rpm 5 bar IMEP. Start of injection is
labeled above each profile and the dashed
symbols represent experimental data while the
solid lines correspond to the simulated cases.
Figure 5 : Influence of Swirl Rate on the Heat
Release Rate Profile at 1500 RPM and 3 bar IMEP.
attributed to the lower engine speed, i.e. lower effective
swirl rate.
Based on this study, it is apparent that swirl rates in the
range of quiescent to that demonstrated on the flow
bench should have minor impact on predicted heat
release rate profiles over the included operating map.
TIMING SWEEP BEHAVIOR Two light load conditions
were investigated to determine if the proposed model
could properly predict the heat release profile response
to an injection timing sweep. As shown in figure 6, the
proposed model does appear to properly follow the
experimental data trends under light-load, medium
speed operating conditions. Quantitatively, the peak
premixed burn rate does respond to ignition timing
changes and also the onset of the mixing controlled burn
also tends to follow the proper trend. The implication of
this small parametric study is that the premixed phase
submodel does properly capture both the mass of fuel
trapped in the shear layer and the turbulent flame
entrainment and burn time scales. Given that the
injection pressure remained essentially constant
throughout these timing sweeps, it is expected that the
global mixing rate profile would remain nearly constant
from point to point and thus the onset of mixing
controlled combustion is phased appropriately with
timing changes.
FULL LOAD BEHAVIOR The heat release profile and
corresponding cylinder pressure for a portion of the full
load curve is shown in figure 7. For the lower two
engine speeds, the model tends to over-predict the
amount of energy released in comparison to the
experimentally determined heat release rate profiles and
thus tends to over-predict the corresponding cylinder
pressure profile. The only feasible explanations for this
discrepancy is that the heat transfer model under-
predicts at lower engine speeds, the fuel consumption
measurements had large experimental errors, or the
experimental heat release analysis did not accurately
capture the burn profile at lower speeds. But, the two
0
2000
4000
6000
8000
10000
12000
14000
16000
-30 -20 -10 0 10 20 30 40 50
Engine Position (degree)
C
y
l
i
n
d
e
r

P
r
e
s
s
u
r
e

(
k
P
A
)
-10
0
10
20
30
40
50
60
70
80
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
Case B
0
2000
4000
6000
8000
10000
12000
14000
16000
-30 -20 -10 0 10 20 30 40 50
Engine P osition (degree)
-10
10
30
50
70
90
110
130
Case A
0
2000
4000
6000
8000
10000
12000
14000
16000
-30 -20 -10 0 10 20 30 40 50
Engine Position (degree)
C
y
l
i
n
d
e
r

P
r
e
s
s
u
r
e

(
k
P
A
)
-10
0
10
20
30
40
50
60
70
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
Case C
0
2000
4000
6000
8000
10000
12000
14000
16000
-30 -20 -10 0 10 20 30 40 50
Engine Position (degree)
C
y
l
i
n
d
e
r

P
r
e
s
s
u
r
e

(
k
P
A
)
-10
0
10
20
30
40
50
60
H
e
a
t

R
e
l
e
a
s
e

R
a
t
e

(
J
/
d
e
g
)
Case D
Figure 7 : Full Load Heat Release Predictions. The dashed symbols are experimental data and the smooth lines
are simulations. Case A:1500 rpm; Case B:2000 rpm; Case C:2500 rpm; Case D:3000 rpm.
higher engine speeds yielded very good agreement
between the simulation and data, and thus more than
likely fuel consumption measurements were not
erroneous. More than likely the discrepancy in released
energy is a combination of poor heat transfer prediction
and discrepancies associated with calculating the heat
release profile from experimental pressure data. Even
with such variances in both the heat release and
pressure profiles, the predictions are acceptable for
conducting engine cycle analysis.
CONCLUSION
A zero-dimensional heat release model has been
developed and compared with data from a small bore,
CIDI engine. The features of this proposed model
include minimal tuning constants, physics-based
premixed and mixing controlled burn submodels, and
real gas behavior for determining the in-cylinder
thermodynamic state. Such tuning constants include a
combustion system-dependent dissipation coefficient, a
temperature dependent laminar flame speed, and a
representative spray cone angle.
Comparison with twenty operating conditions revealed
that the proposed model adequately predicted heat
release rate profiles over a broad timing sweep at both
2000 and 2500 rpm under light load. Similarly, the
proposed model demonstrated the capability to also
predict the full-load heat release profiles within a
tolerance that yielded acceptable cylinder pressure
profiles.
Though a swirl submodel was proposed as part of the
combustion model, minimal data was available over a
large range of swirl numbers to properly assess its
capability. Instead, a brief parametric study was
performed to insure that the initial swirl rate utilized as
an initial condition for the simulations did not fall into a
bandwidth that would substantially modulate the model
tuning process.
ACKNOWLEDGMENTS
The authors wish to thank Ford Motor Company and
FEV Engine Technology for providing the high-speed
cylinder pressure and injection data presented in this
paper. This data was acquired under a joint U.S. Army
TACOM Other Transaction number DAAE07-98-2-0004
with Ford Motor Company sponsored by the National
Automotive Center. The authors also wish to thank Ms.
Renee Bryzik for assisting in high-speed data reduction.
REFERENCES
1. Amsden, A.A., KIVA-3V : A Block-Structured KIVA
Program for Engines with Vertical or Canted
Valves, Los Alamos National Laboratory Report No.
LA-13313-MS, 1997.
2. Amsden, A.A., KIVA-3 : A Kiva Program with Block-
Structured Mesh for Complex Geometries, Los
Alamos National Laboratory, March 1993.
3. Kong, S.C-, Zhiyu, H., and Reitz, R.D., The
Development and Application of a Diesel Ignition
and Combustion Model for Multidimensional Engine
Simulation, SAE Paper 950278, 1995.
4. Jung, D. and Assanis, D., Quasi-Dimensional, Multi-
Zone Turbocharged DI Diesel Engine Simulation for
Performance and Emissions Studies, SAE Paper
2001-01-1246, 2001.
5. Bi, X., Yang, M., Han, S., and Ma, Z., A Multi-zone
Model for Diesel Spray Combustion, SAE Paper
1999-01-0916, 1999.
6. Stiesch, G. and Merker, G., A Phenomenological
Model for Accurate and Time Efficient Prediction of
Heat Release and Exhaust Emissions in Direct-
Injection Diesel Engines, SAE Paper 1999-01-1535,
1999.
7. Bazari, Z., A DI Diesel Combustion and Emission
Predictive Capability for Use in Cycle Simulation,
SAE Paper 920462, 1992.
8. Shahed, S.M., Chiu, W.S., and Yumlu, V.S., A
Preliminary Model for the Formation of Nitric Oxide
in DI Diesel Engines and its Application in
Parametric Studies, SAE Paper 730083, 1973.
9. Lipkea, W.H. and DeJoode, A.D., A Model of a
Direct Injection Diesel Combustion System for Use
in Cycle Simulation and Optimization Studies, SAE
Paper 870573, 1987.
10. Meguerdichian, M. and Watson, N., Prediction of
Mixture Formation and Heat Release in Diesel
Engines, SAE paper 780225, 1978.
11. Dent, J.C. and Mehta, P.S., Phenomenological
Combustion Model for a Quiescent Chamber Diesel
Engine, SAE Paper 811235, 1981.
12. Shipinski, J., Uyehara, O.A., and Myers, P.S.,
Experimental Correlation Between Rate-of-Injection
and Rate-of-Heat-Release in a Diesel Engine,
ASME Paper 68-DGP-11, 1968.
13. Hodgetts, D. and Shroff, H.D., More on the
Formation of Nitric Oxide in a Diesel Engine,
Combustion in Engines, Institution of Mechanical
Engineers, Paper C95/75, p. 129-138, 1975.
14. Hiroyasu, H. and Kadota, T., Models for
Combustion and Formation of Nitric Oxide and Soot
in Direct Injection Diesel Engines, SAE Paper
760129, 1976.
15. Kourenmenos, D.A., Rakopoulos, C.D., and
Karvounis, E., Thermodynamic Analysis of Direct
Injection Diesel Engines By Multi-Zone Modelling,
ASME Winter Annual Meeting, AES vol. 3.3, p. 67-
77, 1987.
16. Li, Q. and Assanis, D.N., A Quasi-Dimensional
Combustion Model for Diesel Engine Simulation,
15
th
Annual Fall Technical Conference of ASME
ICE, vol. 20, p.109-118, 1993.
17. Chmela, F.G. and Orthaber, G.C., Rate of Heat
Release Prediction for Direct Injection Diesel
Engines Based on Purely Mixing Controlled
Combustion, 1999-01-0186, 1999.
18. Watson, N., Pilley, A.D., and Marzouk, M., A
Combustion Correlation for Diesel Engine
Simulation, SAE Paper 800029, 1980.
19. Breuer, C., The Influence of Fuel Properties on the
Heat Release in DI-Diesel Engines, Fuel, vol. 74,
no. 12, p. 1767-71, 1995.
20. Craddock, J.P. and Hussain, M., A Rate of Heat
Release in DI-Diesel Engines, SAE Paper 860083,
1980.
21. Whitehouse, N.D. and Way, R., Rate of Heat
Release in Diesel Engines and its Correlation with
Fuel Injection Data, Symposium on Diesel
Combustion, Institution of Mechanical Engineers,
vol. 184, Pt 3J, p. 17-27, 1969-70.
22. Schihl, P.J., Atreya, A., and Bryzik, W., Simulation
of Combustion in Direct-Injection Low Swirl Heavy-
Duty Type Diesel Engines, SAE Paper 1999-01-
0228, 1999.
23. Magnussen, B.F. and Hjertager, B.H., On
Mathematical Modeling of Turbulent Combustion
with Special Emphasis on Soot Formation and
Combustion, Sixteenth International Symposium on
Combustion, The Combustion Institute, Pittsburgh,
p. 719, 1976.
24. Private communication with U.S. Army.
25. Kanimoto, T., Minagawa, T., and Kobori, S., A Two-
Zone Model Analysis of Heat Release Rate in Diesel
Engines, SAE Paper 972959, 1997.
26. Van Wylen, G., Sonntag, R., and Borgnakke, C.,
Fundamentals of Classical Thermodynamics,
Fourth Edition, John Wiley and Sons, Inc., 1994.
27. Reid, R.C., Prausnitz, J.M., and Poling, B.E., The
Properties of Gases and Liquids, Fourth Edition,
McGraw-Hill Book Company, 1987.
28. Heywood, J.B., Internal Combustion Engine
Fundamentals, McGraw-Hill, Inc., 1988.
29. Mansouri, S.H., Heywood, J.B., and Radhakrishnan,
K., Divided-Chamber Diesel Engines, Part 1: A
Cycle-Simulation which Predicts Performance and
Emissions, SAE Paper 820273, 1982.
30. Poulos, S.G. and Heywood, J.B., The Effect of
Chamber Geometry on Spark-Ignition Engine
Combustion, SAE Paper 830334, 1983.
31. Assanis, D.N. and Heywood, J.B., Development
and Use of a Computer Simulation of the
Turbocompounded Diesel System for Engine
Performance and Component Heat Transfer, SAE
Paper 860329, 1986.
32. Annand, W.J.D., Heat Transfer in the Cylinders of
Reciprocating Internal Combustion Engines,
Institution of Mechanical Engineers, vol. 177, no. 36,
p. 973-990, 1963.
33. Blizzard, N.C. and Keck, J.C., Experimental and
Theoretical Investigation of Turbulent Burning Model
for Internal Combustion Engines, SAE Paper
770647, 1977.
34. Tabaczynski, R., Ferguson, C., and Radhakrishnan,
K., A Turbulent Entrainment Model for Sprak-
Ignition Engine Combustion, SAE Paper 770647,
1977.
35. Dimotakis, P.E., Turbulent Free Shear Layer Mixing
and Combustion, High Speed Flight Propulsion
Systems, edited by S.N.B. Murthy and E.T. Curran,
Progress in Astronautics and Aeronautics, vol. 137,
1991.
36. Spalding, D.B., Mixing and Chemical Reaction in
Steady, Confined Turbulent Flames, Thirteenth
International Symposium on Combustion, The
Combustion Institute, p. 643, 1970.
37. Wakuri, Y., Fujii, M., Amitani, T., and Tsuneya, R.,
Studies of the Penetration of Fuel Spray in a Diesel
Engine, Bulletin of JSME, vol. 3, no. 9, 1960.
38. Dent, J.C., A Basis for the Comparison of Various
Experimental Methods for Studying Spray
Penetration, SAE Paper 710571, 1971.
39. Chikahisa, T. and Murayama, T., Theory and
Experiments on Air-Entrainment in Fuel Sprays and
Their Application to Interpret Diesel Combustion
Processes, SAE Paper 950447.
40. Schihl, P., Bryzik, W., and Atreya, A., Analysis of
Current Spray Penetration Models and Proposal of a
Phenomenological Cone Penetration Model, SAE
Paper 960773, 1996.
41. Naber, J.D. and Siebers, D.L., Effects of Gas
Density and Vaporization on Penetration and
Dispersion of Diesel Sprays, Trans. of SAE, vol.
105, sect. 3, p. 82-111, 1996.
42. Hiroyasu, H. and Arai, M., Fuel Spray Penetration
and Spay Angle in Diesel Engines, Trans. of the
SAE of Japan, 21, p. 5-11, 1980.

You might also like