You are on page 1of 25

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

SAE TECHNICAL
PAPER SERIES

2001-01-1246

Multi-Zone DI Diesel Spray Combustion


Model for Cycle Simulation Studies of
Engine Performance and Emissions
Dohoy Jung and Dennis N. Assanis
The University of Michigan

SAE 2001 World Congress


Detroit, Michigan
March 5-8, 2001
400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A.

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

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

The appearance of this ISSN code at the bottom of this page indicates SAEs consent that copies of the
paper may be made for personal or internal use of specific clients. This consent is given on the condition,
however, that the copier pay a $7.00 per article copy fee through the Copyright Clearance Center, Inc.
Operations Center, 222 Rosewood Drive, Danvers, MA 01923 for copying beyond that permitted by Sections 107 or 108 of the U.S. Copyright Law. This consent does not extend to other kinds of copying such as
copying for general distribution, for advertising or promotional purposes, for creating new collective works,
or for resale.
SAE routinely stocks printed papers for a period of three years following date of publication. Direct your
orders to SAE Customer Sales and Satisfaction Department.
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.

All SAE papers, standards, and selected


books are abstracted and indexed in the
Global Mobility Database

No part of this publication may be reproduced in any form, in an electronic retrieval system or otherwise, without the prior written
permission of the publisher.
ISSN 0148-7191
Copyright 2001 Society of Automotive Engineers, Inc.
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

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

2001-01-1246

Multi-Zone DI Diesel Spray Combustion Model for Cycle


Simulation Studies of Engine Performance and Emissions
Dohoy Jung and Dennis N. Assanis
The University of Michigan

Copyright 2001 Society of Automotive Engineers, Inc.

ABSTRACT
A quasi-dimensional, multi-zone, direct injection (DI)
diesel combustion model has been developed and
implemented in a full cycle simulation of a turbocharged
engine. The combustion model accounts for transient
fuel spray evolution, fuel-air mixing, ignition, combustion
and NO and soot pollutant formation. In the model, the
fuel spray is divided into a number of zones, which are
treated as open systems. While mass and energy
equations are solved for each zone, a simplified
momentum conservation equation is used to calculate
the amount of air entrained into each zone. Details of the
DI spray, combustion model and its implementation into
the cycle simulation of Assanis and Heywood [1] are
described in this paper. The model is validated with
experimental data obtained in a constant volume
chamber and engines. First, predictions of spray
penetration and spray angle are validated against
measurements in a pressurized constant volume
chamber.
Subsequently, predictions of heat release
rate, as well as NO and soot emissions are compared
with experimental data obtained from representative
heavy-duty, turbocharged diesel engines. It is
demonstrated that the model can predict the rate of heat
release and engine performance with high fidelity.
However, additional effort is required to enhance the
fidelity of NO and soot predictions across a wide range
of operating conditions.

heterogeneous combustion has restricted the CIDI


engine penetration in countries with stringent emissions
standards. With the anticipated extreme tightening of
NOx and PM standards by year 2010, intensive research
and development efforts are underway to explore
strategies for meeting those emissions standards, while
retaining the fuel economy benefits of the CIDI engine.

INTRODUCTION

Optimizing the combustion process through options such


as improved combustion chamber design, high pressure
fuel injection systems, and strategic use of EGR
promises significant reductions in engine-out emissions.
Combining such strategies with advanced aftertreatment
and electronic control systems promises that the CIDI
engine has the potential to emerge as the
environmentally friendly, fuel economical powerplant of
the future. Nevertheless, assessing the great number of
available options and their optimum combination is a
very time-intensive task that needs to be addressed
through a smart combination of experimentation and
analysis. Undoubtedly, carefully conducted experiments
can provide relatively precise results for a specific test,
and are therefore needed for ultimate product
certification. However, the cause and effect relationships
implicit in the test results are often hard to interpret, thus
making it difficult to establish strategies that carry-over
from one design iteration to the next through
experimentation alone. On the other hand, modeling and
simulation approaches, although less precise in
predicting the outcome of a specific test, can effectively
isolate one variable at a time and point out trends and
causes. Therefore, a validated CIDI engine simulation
model could be a useful tool for the development of low
emission engines.

Compression ignition, direct injection (CIDI) diesel


engines have been widely used in heavy -duty vehicles
and marine transportation, and are increasingly being
used in light duty vehicles, particularly in Europe and
Japan. The attractiveness of the CIDI engine lies in its
higher fuel economy compared to the spark-ignition
engine due to its lean burn operation, with a higher
compression ratio and without part-load throttling losses.
The power density of the CIDI engine can be competitive
with that of the SI engine with a suitable optimization of
compression ratio and external turbocharging systems,
without knock constraints. Despite its attractive fuel
economy and performance characteristics, the perennial
NOx-particulate matter (PM) tradeoff associated with

Diesel engine simulation models can be classified into


three categories, zero-dimensional, single-zone models,
quasi-dimensional, multi-zone models and multidimensional models. Zero-dimensional, single-zone
models (e.g., [1], [2], [3]) assume that the cylinder
charge is uniform in both composition and temperature,
at all time during the cycle. It has been shown that
calibrated and validated single zone models are capable
of predicting engine performance and fuel economy
accurately and with high computational efficiency.
However, single zone models cannot be used to account
for fuel spray evolution and spatial variation in mixture
composition and temperature, which are essential to
predict exhaust emissions. On the other hand, multidimensional models, like KIVA [4-9] resolve the space of

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

the cylinder on a fine grid, thus providing a formidable


amount
of
special
information.
However,
phenomenological sub-models describing fuel spray
processes are still included in these models, and the
results may vary according to the assumed initial or
boundary conditions. Consequently, the accuracy of the
results cannot be always guaranteed.
Furthermore,
computational time and storage constraints still restrain
these codes from routine use for design purposes.
As an intermediate step between zero-dimensional and
multi-dimensional models, quasi-dimensional, multi-zone
models can be effectively used to model diesel engine
combustion systems. The quasi-dimensional models
combine some of the advantages of zero-dimensional
models and multi-dimensional models. They solve mass,
energy and species equations, but do not explicitly solve
the momentum equation. These models can provide the
spatial information required to predict emission products
and require significantly less computing resources
compared
to
multi-dimensional
models.
Quasidimensional combustion modeling has been an active
area of research since the early heat release studies of
Austen and Lyn [10] and the two-zone combustion
model of Whitehouse and Sareen [11]. Over the years,
numerous models, e.g. [10-32], have been developed to
predict engine combustion with more than one zone.
Within those multi-zone spray combustion models, the
level of detail, fidelity and validation embedded in
individual
sub-models
have
varied
considerably.
Furthermore, only a subset of these models has aimed
at predicting emissions, especially NO and soot.
A number of multi-zone models have tracked the mixing
of gaseous jets with air and subsequent combustion,
without considering the fuel spray dynamics (e.g. [12],
[13], [17], [18], [15], [21], [23]). Kamimoto et al. [17] and
Kobayashi et al. [18] assumed that the fuel injected into
the cylinder vaporizes instantaneously. In other models
proposed by Shahed et al. [12], Chiu et al. [13], and
Lipkea and DeJoode [23], atomization and vaporization
were assumed to be faster than mixing, the spray was
treated as a vapor jet, and no liquid phase was
considered. Kono et al. [21] accounted for variation in air
entrainment rates between the center and outer portion
of the jet by dividing the jet into conical elements and
applying the mass and momentum conservation
equations to each element. However, these approaches
are valid only if the combustion chamber conditions are
near the critical point of the fuel [12]. So, multi-zone jet
models cannot be applied to a wide range of engine
operating conditions.
The most comprehensive class of multi-zone models to
date has followed the framework proposed by Hiroyasu
and his coworkers [14, 19, 20, 26], dividing the spray
into zones in the radial and penetrating directions and
tracking the evolution of the zones over time.
Correlations based on their constant volume vessel
experiments have been widely used by Hiroyasu's group
and others to prescribe the spray cone angle, the spray
break-up length and droplet size distribution, and the
spray tip penetration. The effects of swirl and spray
impingement on walls can be empirically introduced into
the models. It should be noted though that the fuel
injection pressures and thermodynamic conditions near
TDC in modern engines are significantly higher than the
pressures and temperatures under which the Hiroyasu et
al. spray correlations have been developed [33, 34].

Within multi-zone models, combustion models that can


simulate both the premixed and diffusion-controlled
phases are relatively scarce. For instance, Kono et al.
[21] assumed that the combustion rate was only related
to the total amount of air entrainment during premixed
combustion and was independent of the fuel-air mixing
process. Many of the models assume that premixed fuel
and air react only stoichiometrically [e.g. 12, 13, 14, 19,
20, 22, 26, 31]. However, the simple stoichiometric
combustion concept is so sensitive to air entrainment
that ad-hoc calibration coefficients are often applied to
air entrainment rates in order to match the heat release
rate with experimental data. The values of those
empirical coefficients vary very widely among different
references. The stoichiometric combustion concept also
causes overprediction of temperature and NO
emissions, and can produce a sharp drop in heat release
rate during the diffusive combustion phase, as observed
by Kyriakides et al. [22] and Gao and Schreiber [31]. On
the other hand, Bhaskar and Mehta [29] introduced a
combustion model based on the eddy dissipation
concept to handle diffusion-controlled combustion;
however, the concept does not apply to the premixed
combustion phase.
Radiation heat transfer sub-models have not been
included in many multi-zone models (e.g. [14], [16], [19],
[20], [22], [26], [31]). However, estimates of the relative
importance of radiation in cooled diesel engines have
varied between a few and 50 percent of the total heat
transfer [35-43]. Furthermore, models implemented for
the prediction of NO and soot emissions, have often not
captured expected trends, particularly in the prediction of
soot emissions. In an attempt to improve typical two-rate
equation models for soot formation and oxidation (e.g.
[19], [27]), Gao and Schreiber [31] recently employed the
Nagle and Strickland-Constable soot oxidation model, as
implemented by Patterson et al. [8].
In the majority of previous studies, modeling efforts have
focused on the combustion process, so the calculation
covers only the closed part of the cycle or just the
combustion period (e.g. published work by Hiroyasu et
al. [14, 19, 20, 26], Mehta et al. [16, 22, 29], Bazari [24],
Rakopoulos et al. [28, 30], or Gao and Schreiber [31]). In
such cases, initial conditions, such as cylinder pressure,
temperature, and density must be provided as input
data. However, the combustion process controls cycle
temperatures and hence exhaust temperatures. Since
the latter affect the level of boost pressure and thus the
initial conditions for ignition and combustion, full cycle,
multi-zone simulations (such as [25], [27], [32]) are
required in the case of turbocharged engines. In general,
a complete cycle simulation including gas exchange
process is required for practical use as a design tool.
Overall, multi-zone cycle simulations can benefit from
more comprehensive validation exercises, both at the
overall cycle level and the individual sub-model level.
Despite the seemingly large number of available models,
the fidelity of predictions can be improved by embedding
sub-models based on more physically based concepts
and correlations. This exercise would enable application
of the multi-zone models over a wider range of engine
systems and operating conditions. In addition, the
numerical sensitivity of predictions on parameters such
as the degree of zonal or time step resolution needs to
be quantified. The objective of this study is to develop a
more physically-based, quasi-dimensional multi-zone
spray model and implement it into a full cycle diesel
engine simulation so as to predict engine performance,

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

Figure 1. Development of fuel parcels and zones within parcels.


fuel economy and pollutant emissions. The zonal spray
combustion model should be able to:

division of spray into parcels and zones at a certain


instance.

Predict detailed spray evolution with acceptable


fidelity, as evidenced by validation of sub-models.

Incorporate a heat release model that can explicitly


account for both premixed and diffusion-controlled
combustion phases, as observed in measured heat
release profiles.

Predict NOx and soot pollutant emissions with


acceptable fidelity, as evidenced by experimental
measurements.

Include a radiative heat transfer model.

The fuel injected into the chamber is initially assumed to


form a liquid column that travels at a speed equal to the
fuel injection speed until the fuel break-up time elapses.
After that, the injected fuel is distributed within a spray
angle that is unique to each spray parcel and varies from
one time step to another depending on the conditions.
The zone angle, i.e. the injection direction of each zone
is determined by dividing the spray angle with the
number of radial zones. The velocity of each zone is
calculated by temporal differentiation of the correlation
for spray tip penetration. Each zone can be located
relative to the injector hole by tracking the zone angle
and penetration of each zone.

Cover a wide range of engine operating conditions


and engines without losing accuracy.

MULTI-ZONE MODEL ASSUMPTIONS


Figure 1 illustrates the development of fuel parcels and
zones within parcels. Fuel injected into the combustion
chamber according to the fuel injection schedule forms a
parcel during each time step that moves in the spray
axial direction. Each fuel parcel is further divided into
small zones that are distributed in the radial direction.
The zones in each parcel are assumed to contain the
same mass of fuel, however the amount of fuel in the
zones within different parcels may vary according to the
amount of fuel contained within each parcel of fuel. The
mass of fuel in each parcel can be either specified or
calculated by using an empirical correlation based on the
injection and chamber pressures and the injector
geometry. No mixing or passing among zones is
permitted. Individual zones experience their own history
of temperature, pressure and composition. The total
number of zones in the radial direction is fixed
regardless of the amount of fuel injected or the time step
used. However, the total number of zones in the spray
direction equals the number of spray parcels, and is
therefore determined by the injection duration and the
computational time step size. Figure 2 shows the

Following break-up, it is assumed that fuel spray


atomizes to fine droplets, each with a diameter equal to
the Sauter Mean Diameter (SMD). The effect of droplet
size distribution in a spray parcel is neglected. However,
droplet sizes in different parcels may vary according to
the cylinder conditions at the moment of injection. All
calculations related to droplet evaporation are based on
SMD.
The air entrainment rate depends on the physical
position of each zone, with centerline zones receiving
less and edged zones receiving more air. The amount of
entrained air is calculated based on conservation of
momentum applied to each zone. It is assumed that the
momentum of the zone at any instant is equal to the
i-Radial direction, k-Injection direction
Zone Angles

(3,1)
Spray Angle
(i,k)

(2,1)

(2,2)
(1,2)

Parcel :

(1,1)

(3)
(2)
(1)

Spray Penetration

Figure 2. Division of spray into parcels and zones at a


certain instance.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

momentum imparted in the zone upon nozzle exit. Since


the mass of fuel and injection velocity of each zone is
initially determined and the velocity of the zone can be
subsequently calculated, the amount of air entrained is
obtained by the momentum conservation equation.
It is assumed that fuel droplets begin to evaporate
immediately after break-up occurs. Both heat and mass
transfer for a single evaporating droplet are considered
in order to compute instantaneous droplet temperature,
rate of evaporation and droplet diameter.
Combustion is assumed to start individually in each zone
after the lapse of the ignition delay period. The ignition
delay is measured from the start of injection and is
calculated based on zonal temperature and pressure.
During the ignition delay period, some of the injected fuel
is evaporated and mixed with air, forming a combustible
mixture. In the early stage, combustion occurs under
premixed conditions. Premixed combustion is assumed
to occur until the amount of fuel evaporated at the end of
the ignition delay period has been consumed. When the
entire initial fuel vapor has been consumed, combustion
is assumed to be controlled by diffusion of air into fuel
zones.
NO and soot are calculated depending on
pressure, zonal temperature and composition.

SUBMODELS FOR MULTI-ZONE COMBUSTION


MODEL
FUEL INJECTION
The timing and rate of fuel injection into the chamber
affect
the
spray
dynamics
and
combustion
characteristics. If the pressure upstream of the injector
nozzle can be estimated or measured, and assuming the
flow
through
each
nozzle
is
quasi-steady,
incompressible, and one dimensional, the mass flow rate
& i , is given by
of fuel injected through the nozzle, m
Heywood [44]:

m& i = CD An 2 l P

(1)

where CD is a discharge coefficient; An is the nozzle hole


area; l is the density of liquid fuel; P is pressure drop
& i = l An ui , the fuel
across the injector nozzle. Since m
injection velocity at the nozzle tip, ui, can be expressed
as:

u i = C D 2P / l

(2)

FUEL SPRAY DYNAMICS


Spray Penetration
Unlike multi-dimensional models in which the entire
combustion chamber is taken as the computational
domain, only the fuel spray is divided into a number of
zones in the multi-zone model. The control volume of
each zone is treated as an open system, and mass and
energy equations are solved for each zone. Instead of
solving the full momentum equation, which is one of the
dominant reasons for the computational inefficiency of
multi-dimensional models, the multi-zone model

depends on empirical correlations to describe spray


penetration over time.
Utilizing this method, quasidimensional models can offer the fastest and least
expensive means of generating the spatial information
required to predict emission products. Since the multizone model depends on an empirical correlation for
spray evolution, the fidelity of the spray penetration
model is crucial for accuracy.
Hiroyasu and Arai [45] proposed the following
correlations for spray penetration before and after
breakup.
(a) Before breakup, 0<t<tb(I)
0. 5

2P
t
S = 0.39
l

(3)

(b) After breakup, tb(I)t

S = 2.95
a

0 .25

( d n t ) 0 .5

(4)

where breakup time, tb, is

t b = 28.65

l d n
( a P) 0 .5

(5)

and S is spray tip penetration; P is pressure drop


through the nozzle hole; is density; dn is nozzle hole
diameter; and subscript l and a denote liquid fuel and
ambient gas respectively. These correlations have been
widely used in multi-zone models by independent
researchers [19, 22, 24, 31].
The value used in equations (3) and (4) for the discharge
coefficient of the nozzle was 0.39. However, the
discharge coefficient depends on the nozzle geometry,
and discharge coefficients of injector nozzles in modern
diesel engines usually range between 0.6 and 0.8. To
generalize the Hiroyasu correlations so as to handle
nozzles with different discharge coefficients other than
0.39, their correlation is modified as follows:
(c) Before breakup, 0<t<tb(I)
0 .5

2P
t
S = C D
l

(6)

(d) After breakup, tb(I)t

S = 2.95
a

0 .25

( d n t ) 0 .5

(4)

where breakup time, tb, is

t b = 4.351

l d n
C D ( a P) 0. 5
2

(7)

A detailed derivation of the above correlations is given in


the Appendix.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

AIR ENTRAINMENT

Spray Angle
The two most commonly used correlations for spray
angle by Hiroyasu and Arai [45] and Reitz and Bracco
[46] have been implemented into the multi-zone model.
The correlation by Hiroyasu and Arai is:

Pd 2
= 0.025 a 2 n
a

0 .25

(8)

where a is the viscosity of ambient gas. The Reitz and


Braccos correlation, which appears to yield predictions
in better agreement with experimental data (see
discussion under Spray Validation), is:

1
tan( ) = 4 a
A l

The air entrainment rate into a given zone is controlled


by the conservation of momentum applied to that zone.
It is assumed that the initial momentum imparted in a
zone upon nozzle exit is equal to the momentum of the
zone at any subsequent distance traveled, i.e.

1/2

3
6

m f ui = ( m f + ma )

dt
ma = m f ui
1
dS

(10)

m f ui d 2 S
dS 2 dt 2
dt

m& a =

and ln is nozzle hole length.

Following breakup, all droplets in a zone are assumed to


have the same initial diameter, equal to the Sauter Mean
Diameter (SMD), thus neglecting droplet size distribution
and the details of the atomization process. The following
equations [47], which are commonly used in the multizone models, give the SMD, d32, in the zone located
along the spray centerline, with the radial SMD
distributions assumed to be normal. However, allowance
is made for droplet size variations from one fuel parcel to
the next depending on the operating condition.
HS
d 32LS d 32
d 32

= MAX
,
dn
dn dn

(11)

where

= 0.38 Re i0 .25 We i 0. 32 l
dn
a

0 .54

0 .37

l

a

0. 18

0 .47

Following the break-up period, it is postulated that liquid


fuel in zones is atomized into fine droplets of size equal
to the SMD. Each individual zone has its own initial
SMD. Together with the amount of fuel contained in the
zone, the initial droplet size gives the number of
droplets, N, contained within a zone:

N=

ml, i

(19)

3
d 32
l / 6

where ml,i is fuel mass injected into a zone. The number


of droplets in a zone is not changed. However, the
diameter of all droplets within a zone decreases during
the process due to evaporation. At any instance, the
mean diameter of droplets, dl, in a zone is given by:

[(

d l = 6 ml , i dm fg Nl

(12)

1/ 3

(20)

where mf g is the mass of gaseous fuel.


(13)

Re i =

l u i d n
l

(14)

Wei =

u d n l

(15)

2
i

(18)

FUEL EVAPORATION

Droplet Diameter after Break-Up

HS
32

(17)

By differentiating equation (17) with respect to time, the


air entrainment rate can be obtained as:

l
A = 3.0 + 0.28 n
dn

d 32LS

= 4.12 Re i0 .12 We i 0. 75 l
dn
a

(16)

where mf is the mass of fuel in a zone and ma is the


mass of air in the zone. Therefore, the amount of
entrained air of a zone is proportional to the decrement
in the zone velocity. Equation (16) can be rewritten as:

(9)

where

dS
dt

l is the viscosity of liquid and is the surface tension.


By using the above equations and assuming normal
distributions, initial diameter of liquid fuel droplets right
after the breakup time in each zone is calculated.
Therefore, the number of drops in each zone can be
determined knowing the SMD and the mass of fuel
injected.

The rate of evaporation is determined by combining


equations governing mass diffusion and heat transfer.
The rate of diffusion of vapor away from the droplet is
given by Borman and Johnson [48].

dm fg
dm
= l
dt
dt
= d l NDv Sh

Pt
Pt
ln

RvTm Pt Pv, surf

(21)

where

Tm =

Ta + Tl
2

(22)

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

and Dv is the mass diffusivity; Sh is the Sherwood


number; Pt is the total pressure; Rv is the gas constant;
Ta is the ambient gas temperature; Tl is the bulk
temperature of liquid droplet. The partial pressure of
vapor at the droplet surface, Pv,surf , is equal to the
saturation pressure corresponding to the temperature of
the droplet.
Convective heat transfer from the hot gas to liquid fuel
droplets in a zone, q, is modeled following the
methodology of Gosman and Johns [49].

z
q = dl Nkm (Tg Tl )Nu z

e 1

(23)

where Tg is the bulk gas temperature; k m is mean


thermal conductivity that is defined as:

k m = (1 Pv , surf / 2 Pt ) k a + ( Pv , surf / 2 Pt ) kv

(24)

where ka is the thermal conductivity of the ambient gas


and k v is the thermal conductivity of vaporized fuel.
Expressions for the Nusselt and Sherwood numbers are
obtained from the following correlations proposed by
Ranz and Marshall [50] for a single droplet.

Nu = 2 + 0.6 Re d
Sh = 2 + 0.6 Re d

12

12

Pr 1 3

(25)

13

(26)

Sc

The Reynolds, Prandtl and Schmidt numbers are defined


as follows:

m ud l
m
C
Pr = Pm m
km
m
Sc =
m Dv
Re d =

(27)

(28)

(29)

Pt M m
R Tm
/ 2 Pt ) M a + ( Pv , surf / 2Pt ) M v

M m = (1 Pv , surf

m = (1 Pv , surf / 2 Pt ) a + ( Pv , surf / 2 Pt ) v

dm fg

dt
d l Nk m Nu

(34)

The resulting rate of liquid fuel temperature rise is given


by the energy balance as:

dTl
1
=
dt
ml C P l

dm fg

dt

(35)

where is the specific heat capacity of liquid fuel.

IGNITION DELAY
Ignition delay time can be calculated as the difference
between the time at which combustion starts, tign, and
the time at which injection starts, tinj. In general, ignition
delay is a complicated function of mixture temperature,
pressure, equivalence ratio, and fuel properties. For the
purposes of diesel combustion simulations, simplified
Arrhenius expressions of the form

2100
= 3.45 P 1 .02 exp

(36)

have been used [51], where P is gas pressure, and T is


gas temperature. This expression is adopted here. Note
however that pressure and temperature change
considerably during the ignition delay period due to the
compression resulting from piston motion. To account for
these changing conditions, tign can be obtained by
integrating the reciprocals of instantaneous estimates of
the ignition delay, based on equation (36), until the
following relationship is satisfied [1].

t ign

t inj

1
dt = 1

(37)

HEAT RELEASE DURING PREMIXED AND


MIXING-CONTROLLED COMBUSTION PHASES

The mean values in the above expressions are:

m =

z=

C Pv

(30)
(31)
(32)

C P m = (1 Pv , surf / 2 Pt )C P a + ( Pv, surf / 2 Pt )C P v (33)


where Cp is the specific heat; subscript a denotes
ambient gas, and subscript v denotes fuel vapor. Dv , a,
v , ka, kv , Cpa, and Cpv are evaluated at the temperature
Tm while Pv , l, and Cpl are evaluated at Tl. In the present
application, the fuel is assumed to be a single pure
substance having approximately the same molecular
weight as diesel fuel, namely, n-dodecane (C12H26),
whose properties are obtained from Borman and
Johnson. To correct for the effects of boundary layer
thickening due to mass transfer, the heat transfer
z
coefficient is multiplied by the factor z/(e -1) in equation
(23). z is given by:

During the ignition delay period, some of the injected fuel


is evaporated and mixed with air, forming a combustible
mixture. Following ignition, the first phase of combustion
occurs under premixed conditions at a rate RRp given by
the following Arrhenius type kinetic equation [52]:

1200
2
Vz
RR p = B1 mix
x fv x 5ox exp
T

(38)

B1 is frequency factor; mix is density of the mixture; xf v is


the mass fraction of fuel vapor and xox denotes the mass
fraction of oxygen; Tz is temperature of the zone; Vz is
volume of the zone. Equation (38) is assumed to be valid
until the amount of fuel evaporated during the ignition
delay period has been consumed.
After the entire initial fuel vapor has been consumed,
combustion is assumed to proceed to the mixingcontrolled and late combustion phases. Accordingly, the
combustion model is formulated so that it can handle
both mixing and kinetics-limited combustion. Under
normal conditions, the combustion during the second

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

phase is mixing-limited, so it proceeds at a rate limited


by the availability of fuel vapor and entrained air. In this
situation, the rate at which the combustion kinetics
proceeds is considered to be so rapid that it provides no
additional limit on the combustion process. However,
when the gas temperature becomes low, the combustion
kinetics slow down exponentially, and at some point
become the limiting factor. This can occur at the latter
parts of the expansion process or at cold start. The
kinetics can also slow down when the mixture is very
lean. It is therefore proposed that the combustion rate for
the mixing-controlled and late combustion phases, RRm ,
are governed by the following expression:

RR m = B2 m fv

2500
Pox 02 .5

P exp
P
Tz

(39)

where B2 is a constant; mf v is the mass of fuel vapor; Pox


is the partial pressure of oxygen; P is the total pressure.
If the combustion is mixing-limited, i.e. the amount of fuel
available in the zone is less than the one prescribed by
equation (39), only the available fuel is burned. If the
combustion is kinetically limited, the fuel is burned at the
rate determined by equation (39).

[ ] denotes species concentration in kmoles per cubic


meter and subscript e means equilibrium. The NO
concentration in equation (46) can be converted to mass
fraction as:

dX NO 2(M NO C .V . )R1 1 ([ NO] [ NO]e )


=
dt
1 + ([ NO] [ NO] e ) R1 ( R2 + R3 )

(50)

where XNO is the mass fraction of NO; MNO is the


molecular weight of NO; and C.V. is the density of the
control volume. Zones in the quasi-dimensional model
are not considered as a closed system, but an open
system, since the air entrainment into each zone is an
incoming flow to the zone.
Accordingly, the mass
fraction of NO of each zone cannot be calculated simply
by applying the equation for extended Zeldovichs
mechanism, equation (50). The effect of air entrainment
should be considered. The derivative of the NO mass
fraction within a zone with respect to time is:

DX NO
D m NO

=
Dt
Dt mtot
=

POLLUTANT EMISSIONS

dX NO
dt

X dmtot
NO
mtot dt

(51)

Since the total mass of a zone is changed only by air


entrainment, the last term in equation (51) is the same
as the air entrainment rate that is given in equation (18).
So, equation (51) becomes:

Nitric Oxide Formation


While nitric oxide (NO) and nitrogen dioxide (NO2) are
usually grouped together as NOx emissions, NO is
predominant in diesel engines [44]. Therefore, only NO
formation is considered in the present study. The
principal reactions governing the formation of NO from
molecular nitrogen and its destruction are [53]:
k1
N + NO
N2 +O
k2
O + NO O2 + N

(40)
(41)

k3
N + OH
NO + H

(42)

where forward rate constants, k1, k 2, and k3 are:

k 1 = 1.6 1010

(52)

and by substituting equation (50) into equation (52), the


time derivative of the mass fraction of NO for an open
system is obtained:

DX NO 2(M NO C.V . )R1 1 ([ NO] [ NO]e )


=
Dt
1 + ([ NO] [ NO]e ) R1 (R2 + R3 )
m&
(53)
a X NO
mtot
2

(43)

19500
k 2 = 1.5 106 T exp

k 3 = 4.1 1010

Soot Formation and Oxidation


(44)
(45)

This is often called the extended Zeldovichs


mechanism. Zeldovich was the first to suggest the
importance of reactions (40) and (41). Lavoie et al. [53]
added reaction (43) to the mechanism. The rate of
change of NO concentration is expressed as [44]:

DX NO dX NO m& a
=

X
Dt
dt
mtot NO

d [ NO]
2 R1 1 ([ NO] [ NO]e )
=
dt
1 + ([ NO] [ NO] e ) R1 ( R2 + R3 )
2

(46)

Soot forms in the rich unburned-fuel-containing core of


the fuel sprays, within the flame region, where the fuel
vapor is heated by mixing with hot burned gases. Soot
then oxidizes in the flame zone when it contacts
unburned oxygen. Therefore, the concentration of soot
in the exhaust is governed by the formation and
oxidation of soot during the engine cycle, i.e.

dms dmsf dmso


=

dt
dt
dt

(54)

where m is mass; subscript s, sf, and so denote soot


emitted, soot formed, and soot oxidized, respectively.

where

R1 = k1 [ NO]e [ N ]e
R2 = k 2 [ NO] e [ O] e
R3 = k 3 [ N ] e [OH ]e

(47)
(48)
(49)

The general fact that the net soot formation rate is


primarily affected by pressure, temperature and
equivalence ratio has been fairly well established.
However, the details of the mechanism leading to soot
formation are not known. Consequently, semi-empirical,
two-rate equation models have been used to describe

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

the soot dynamics. In particular, the soot formation


model proposed by Hiroyasu et al. [19] is used in many
multi-zone models. The formation rate is calculated by
assuming a first-order reaction of vaporized fuel, mf g, as:

Esf
= A f m fg P 0 .5 exp
dt
RT

dmsf

(55)

The soot oxidation is predicted by assuming a secondorder reaction between soot, ms , and oxygen.

dmso
P
Eso
= Ao m s ox P1 .8 exp

dt
P
RT
4

(56)

where Esf =1.2510 kcal/kmol, Eso=1.4010 kcal/kmol.


Af and Ao are constants that are determined by matching
the calculated smoke with the measured soot in the
exhaust gas.
Another soot oxidation model investigated in this study is
the Nagle and Strickland-Constable [54] oxidation
model. The NSC oxidation model is based on oxidation
experiments of carbon graphite in an O2 environment
over a range of partial pressure. In this model, carbon
oxidation occurs by two mechanisms whose rates
depend on the surface chemistry involving more reactive
A sites and less reactive B sites. The chemical reactions
are:

A + O2 A + 2CO
B + O2 A + 2CO
A B

(57)
(58)
(59)

The NSC soot oxidation rate implemented in equation


(54) is given by:

dmso
MC
=
mw
dt
s d s s

x + K B Pox (1 x )

Pox
Pox + ( KT / K B )

(61)

(62)

The rate constants used in the NSC oxidation model are


given in Table 1.
Table 1. Rate constants for NSC soot oxidation model
Rate Constant
units
KA=20 exp(-15100/T)
-3

KB=4.4610 exp(-7640/T)
5
KT=1.5110 exp(-48800/T)
KZ =20 exp(-15100/T)

CONSERVATION OF MASS
The rate of change of the total mass in any open system
is equal to the sum of the mass flow rates into and out of
the system.

m& i = m& i , j
j

(63)

where i denotes i-th zone; and j denotes mass flows to


and from the zone. During the exhaust, intake, and
compression processes only the air zone, which is
considered as an open system, exists inside the
cylinder. Gas flows through valves are mass flows into
and out of the system. Once the fuel is injected, spray
zones are generated insider the cylinder. Air entrainment
is the mass flow for each spray zone and air zone. In
particular, conservation of the fuel species can be
expressed as:

m& f , i =

m&

f , i, j

(64)

where Pox is the oxygen partial pressure in atm. The


proportion, x, of A sites is given by:

x=

In this section, the implementation of the multi-zone


combustion model into the thermodynamic cycle
simulation developed by Assanis and Heywood [1] is
presented. Detailed information about modeling of
reciprocating engine processes, including gas exchange
during the intake and exhaust strokes, turbulence,
convective heat transfer, and radiative heat transfer, as
well as modeling of external engine system components,
such as turbo machinery and intercooler can be found in
Assanis and Heywood [1].

(60)

where MC is the carbon molecular weight (12 g/mole), s


3
is the soot density
(2.0 g/cm ), and ds is the soot
-9
diameter (4.510 m). The term w in equation (60) is the
net reaction rate of reactions (57), (58), and (59) and is
defined as:

K A Pox
w =
1 + K Z Pox

COMBUSTION MODEL IMPLEMENTATION IN


THERMODYNAMIC CYCLE SIMULATION
FRAMEWORK

g-C/cm s atm
2
g-C/cm s atm
2
g-C/cm s
-1
atm

where mf,i denotes the fuel content in the i-th zone as an


open system (includes fuel added by evaporation from
the liquid fuel droplets and fuel in the form of combustion
products). Defining the fuel fraction, Fi, in the system as
Fi = mf,i /mi where mi is the total mass in the i-th zone,
equation (64) can be rewritten as:

d ( mi Fi )
& i , j Fi , j
= m
dt
j

(65)

where Fi,j denotes the fuel fraction of the mass flow


entering or leaving the open system. Differentiating the
&i
left-hand side of equation (65) and substituting for m
from equation (63) results in a differential equation for
the change in the fuel fraction of the open system, i.e.

F&i = ( m& i , j / mi )( Fij Fi )

(66)

An average fuel-air equivalence ratio, , for the contents


of the open system can be defined as:

m f ma
( m f ma ) s

(67)

where ma is the mass of air in the open system and the


subscript s denotes the stoichiometric fuel to air ratio.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

Expressing the equivalence ratio in terms of the fuel


fraction and differentiating with respect to time, we
obtain an equation for the rate of change of the
equivalence ratio of the zone, i.e.

&i =

1
F&i
( m f , i / ma , i ) (1 Fi ) 2

(68)

Q& w = Q& c + Q& r

(76)

Convective heat loss from the gas to the wall is based


on the turbulent flow model of Assanis and Heywood [1].
In addition, details of radiative heat loss can be also
found in Assanis and Heywood [1]. Heat transfer to the
liquid fuel droplets of each zone can be computed using
equation (23).

z
q i = d l , i N i k m , i (Tg , i Tl ,i )Nu i z i i
e 1

CONSERVATION OF ENERGY
The first law of thermodynamics, supplemented with a
heat transfer model, can be applied to yield cylinder
pressure and temperatures of the air zone and spray
zones. Treating the contents of each zone as an ideal
gas, with thermodynamic properties that are functions of
temperature, pressure, and fuel-air equivalence ratio,
Assanis and Heywood [1] showed that:

V& 1 & 1 & m&



P& =
T
+ (69)
/ P V T

m
and

(77)

Accordingly, the total heat transfer to the fuel liquid


droplets is given by:

Q& fg =

(78)

i = zones

The energy equation for each zone as an open


thermodynamic system can be expressed from equation
(70) as:

mi CPiT&i = m& i , j hi , j + Q& i + (Vi CTi ) P& mi Ci&i m& i hi


j

(79)

mC P T& = m& j h j + Q& + (V CT ) P& mC & m& h

or

(70)
where the first term in equation (70) is the net rate of
influx of enthalpy; Q& is the total heat transfer to the
system; and the dot denotes differentiation with respect
to time. Heat release due to combustion is treated as the
difference of the enthalpy of products and the enthalpy
of reactants.
By substituting for P& from equation (69), Assanis and
Heywood obtained an equation for the rate of change of
temperature that does not explicitly depend on the rate
of change of pressure of the system, i.e.

B m&
h
V& C
1
m& j h j + Q&
T& = 1 & +

A m
B V B
Bm j

(71)
where

A = CP +
B=

( / T ) 1
CT
( / P)

( / ) 1
CT
( / P)

(80)
Since the equivalence ratio of the air zone does not
change after the intake stroke, the rate of change of the
air zone temperature can be derived from equation (80)
as:

T&a = m& a , j ha , j + Q& a + (Va CTa ) P&i m& a ha ma CPa


j

(81)
The total heat transfer rate given by equation (75) is
distributed to each zone by weighing its components
with the zonal mass and temperature, as follows:

Q& i = qi

mi Ti
Q& c
m a Ta + m i Ti

mi Ti
Q& r
m
T
ii

i = zones

mi Ti

m T

i = zones

i i

(82)

1
(1 CT )
( / P)

C = C +

(72)

& i , j hi , j + Q& i + (Vi CTi ) P& miCi&i m


& i hi miC Pi
T&i = m
j

(73)

SPRAY MODEL VALIDATION IN A CONSTANT


VOLUME CHAMBER
(74)

The total heat transfer is the sum of heat transfer to the


cylinder wall via convection and radiation and heat
transfer to the liquid fuel droplets during the evaporation
process. Therefore, the heat transfer rate term, Q& , in
equation (71) is expressed as:

Q& = Q& w Q& fg


Heat transfer rate to the wall, Q& w , is given as:

(75)

SPRAY PENETRATION
To
validate
the
spray
penetration
correlation,
experimental data collected by Dan et al. [34] were
compared with our predictions. The injector specification,
injection conditions and ambient conditions of the
experiment are summarized in Table 2. The injector
nozzle has a mini-sac volume design for high injection
pressure. The injection pressure was varied from 55

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

MPa to 120 MPa. The ambient gas was nitrogen and the
fuel was n-tridecane (C13H28), with hydrodynamic
properties given in Table 3, as reported by Dan et al.
[34].
Table 2. Experimental conditions
Parameter
Hole diameter (mm)
Hole length (mm)
Number of holes (-)
Discharge coefficient of the hole (-)
Injection pressure (MPa)
Ambient pressure (MPa)
Ambient temperature (K)
3
Ambient density (kg/m )
-6
Ambient viscosity (x10 Pa s)

Value /Spec.
0.2
1.1
1
0.66
55, 77, 99, 120
1.5
293
17.3
17.5

Table 3. Hydrodynamic properties of fuel


Property
n-tridecane
3
Density (kg/m )
756.2
-6
Kinetic viscosity (m2/s) 2.48x10
-3
Surface tension (N/m)
26.13 x 10

(a) 120 MPa

(T=293K)
Diesel light oil
831-951
-6
2.5 x10
-3
29-30 x 10

Figure 3 compares predictions with measurements of the


spray tip penetration as a function of time from start of
injection and injection pressure. In Figure 3a, the
measured spray penetration for an injection pressure of
120 MPa is compared with computed data from both a
multi-zone model and a multi-dimensional model. The
latter is based on KIVA-3V, modified by Nishimura and
Assanis [55] to incorporate a model for primary diesel
fuel atomization based on cavitation bubble collapse
energy. Both models predict measured spray tip
penetration with good accuracy. The multi-zone model,
though, results in very substantial computational savings
(on the order of 50-100 times) compared to the multidimensional model. In Figures 3b through Figure 3d,
measured spray tip penetrations for injection pressures
of 99, 77, and 55 MPa, respectively, are compared with
calculated data from the multi-zone model. It can be
concluded that spray tip penetrations predicted by the
multi-zone model exhibit good overall agreement with
experimental data over a wide range of injection
pressures. One notable discrepancy is that the multizone model tends to overpredict spray tip penetration in
the very early stages of fuel injection.

(b) 99 MPA

(c) 77 MPa

SPRAY ANGLE VALIDATION


Two empirical correlations for the spray angle (Hiroyasu
and Arai [45], Reitz and Bracco [46]) have been
implemented into the multi-zone model. Spray angles
predicted with those correlations at different injection
pressures are compared with the experimental data from
Dan [56] in Figure 4. As can be seen in Figure 4, the
Hiroyasu and Arai correlation overpredicts measured
spray angle, and the predicted spray angle increases as
the injection pressure increases. However, the
experimental data shows that the measured spray angle
does not vary significantly as injection pressure
increases. On the other hand, the Reitz and Bracco
correlation underpredicts spray angle, but the
differences between their predicted spray angles and
measured data are smaller than those from Hiroyasu

(d) 55 MPa

Figure 3. Comparison of predicted and measured


temporal spray tip penetration for a range of injection
pressures.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

and Arais correlation. Furthermore, the Ritz and Bracco


correlation is not a function of injection pressure. The
lack of pressure dependence is also borne out by the
experimental trends observed by Dan. Figure 5
compares the overall spray shapes simulated by the
multi-zone model, using each of the alternative spray
angle correlations, with the photographs taken by Dan.
Clearly, the spray shape simulated with Hiroyasu and
Arais correlation is much wider than the actual shape.
Therefore the Reitz and Braccos correlation is employed
for the remainder of the simulation studies reported here.

Figure 4. Comparison of spray angle predicted using


two alternative spray angle correlations against
experimental data [56] for a range of injection pressures.

need to be calibrated against experiments. The purpose


of this effort is to explore the range of engine speed,
load and injection timing conditions over which the multizone model predictions remain valid, following only an
initial calibration. An additional objective is to explore the
sensitivity of the model to discretization parameters,
such as the zonal and time step resolution, so as to
optimize its numerical performance and robustness.

VALIDATION AGAINST MULTI-CYLINDER


ENGINE EXPERIMENTS
To calibrate the multi-zone spray combustion model and
subsequently assess its fidelity in predicting cycleresolved and cycle-integrated performance parameters,
measurements were taken on a representative, modern,
electronically controlled diesel engine commonly found
in heavy -duty trucks. Primary specifications of the fourstroke, in-line, six cylinder, turbocharged, intercooled,
water-cooled, direct injection diesel engine are listed in
Table 4. The relatively quiescent combustion chamber
employs a shallow Mexican hat bowl-in piston and very
high injection pressures, delivered by unit injectors. The
fuel injection timing and duration are electrically
controlled. More details on the experimental setup at the
Automotive Research Center of The University of
Michigan can be found in Filipi et al. [57], Assanis et al.
[58], and Fiveland [59]. Measured data, which included
crank-angle resolved cylinder pressure profiles and
supporting low speed data (air and fuel flow rates,
torque), will be compared with simulation predictions
over a wide range of speeds and loads.
Table 4. Heavy duty diesel engine specification
Displacement volume (liter)
12.7
Bore (mm)
130
Stroke (mm)
160
Con. Rod length (mm)
269.3
Compression ratio (-)
15 : 1
Rated speed/power (rpm/kW)
2100/350
Calibration and Sub-Model Behavior

(a)

(b)

(c)

Figure 5. Comparison of spray shapes from the


(a) multi-zone model with Reitz and Braccos correlation
(b) experiments [56], and (c) multi-zone model with
Hiroyasu and Arais correlation.

MULTI-ZONE MODEL VAL IDATION IN REAL


ENGINES
In this section, multi-zone model predictions of engine
performance and emissions have been compared
against corresponding measurements on a modern
multi-cylinder engine and a single-cylinder test engine.
Critical sub-models, such as for the heat release and
emissions processes, contain empirical constants that

A single operation point (1800rpm, 50% load) is selected


for calibration of constants, B1 and B2 in the two-phase
heat release model, equation (38) and equation (39).
Those two constants were calibrated to match
predictions with measurements of the cylinder pressure
and the apparent heat release rate profiles. The values
selected for B1 and B2 were 1.5 and 800, respectively.
The rest of the multi-zone model parameters were set
according to baseline values reported in the literature.
Figure 6 shows the mass flow versus crank angle
profiles corresponding to the filling and emptying events
of the cylinder. The intake valve flow rate essentially
follows the piston motion, and exhibits a gradual rate of
acceleration and deceleration. In contrast, the exhaust
flow rate changes more rapidly and exhibits two peaks.
The first peak, corresponding to the blowdown process
of the cylinder, is caused by the high cylinder pressures
and relatively o
l w exhaust pressures when the exhaust
valve opens. The second peak, lower in magnitude, is
due to the upward piston motion during the exhaust
stroke. Near the closing of the intake and the exhaust
valves, backflow (shown as negative mass flow rates)
can occur when the cylinder pressure exceeds the intake
manifold pressure. This is an undesirable scavenging

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

effect, since it reduces the amount of fresh air that is


trapped in the cylinder.
Figure 7 shows predicted and measured cylinder
pressure traces at the operating condition selected for
model calibration. The predicted cylinder pressure trace
falls in very good agreement with the experimental data.
Note that predicted peak pressure is 10.62 MPa at 369
degrees CA, and measured peak pressure is 11.0 MPa
at 368.75 degrees CA. Therefore, both the magnitude
and timing of occurrence of the peak pressure are
precisely predicted by the model. The observed cylinder
pressure profiles reflect the effects of in-cylinder heat
release, heat transfer to the cylinder surfaces and work
transfers. Even though predictions of the gross heat
release rate are directly available from the model, the
apparent heat release rates have been calculated so as
to compare them with experimental data. The equation
for the apparent heat release rate, determined from
pressure data, is given by [44]:

dQn

dV
1
dP
=
P
+
V
dt
1 dt 1 dt

(83)

The ratio of specific heats is not necessarily constant


through the compression, combustion, or expansion
process. Here, the ratio is calculated by a relatively
simple correlation for the purpose of comparing
measured and predicted heat release rates, as follows:

= 1.338 6.0 10 T + 1 10 T
5

(84)

The fuel injection rate that is prescribed to the multi-zone


model is obtained using equation (1) for the given
injection pressure and cylinder pressure. In order to
measure the injection pressure, a fuel cam rocker arm
was instrumented with a temperature-compensated, 90
degree Rosette strain gauge. This technique was
chosen since the unit injector design makes internal
instrumentation extremely difficult. The strain gage with
its half-bridge circuit was mounted at the center-top
location on the topside of the fuel rocker arm, as detailed
by Filipi et al. [57]. By deducing the inertial force and the
force to compress the injector follower spring from the
force measured by the strain gage, the pressure force on
the plunger was obtained. For a known plunger
diameter, this force balance directly yields the injection
pressure. The dynamic start of fuel injection was
determined from the measured injection pressure by the
skip-fire technique [58].
Figure 8 compares apparent heat release rates
computed from predicted and measured pressure traces.
Also shown is the fuel injection profile. Note that the heat
release rate decreases from the start of injection to the
start of combustion (ignition delay period) because of the
fuel evaporation occurring during this period. Figure 8
exhibits the double-peak shape that characterizes the
direct injection diesel combustion. The first peak due to
premixed combustion strongly depends on the amount of
fuel that is prepared for combustion during the ignition
delay period. The second peak due to diffusion
combustion is controlled by the fuel-air mixing rate.
Diffusion combustion continues until combustion is
completed. As can be seen in Figure 8, the shape,
timing, and magnitude of both heat release peaks
predicted by the model compare well with experimental
data. This effectively validates that the spray dynamics,
fuel droplet evaporation, fuel-air mixing, ignition delay,
and combustion sub-models are properly working.

Figure 6. Predicted mass flow rates through the


intake
(top)
and
exhaust
(bottom)
valves,
corresponding to the filling and emptying processes of
the cylinder.

Figure 7. Comparison of predicted and measured


cylinder pressure traces at 1800 rpm, 50% load.

Figure 8. Comparison of apparent heat release rates


computed from predicted and measured pressure
traces at 1800 rpm, 50% load. Also shown is the rate
of fuel injection.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

Figure 9 shows profiles of air entrainment into the spray


zones and cumulative air consumption due to
combustion. Similarly, Figure 10 illustrates the
preparation for and usage of fuel during the expansion
process. The cumulative mass of injected fuel increases
throughout the injection period. Injected fuel evaporates
very fast right after breakup occurs. Since the
combustion is mainly mixing-controlled after the
premixed combustion phase, the combustion rate curve
during the diffusion combustion phase in Figure 8 is
quite similar to the shape of the air entrainment rate
curve of Figure 9.
Figure 11 compares the predicted temperatures of spray
tip, tail, and air zones with the global gas temperature.
The temperature of the spray tip zone increases sharply
right after the ignition delay period. The faster rate of
temperature rise at the spray tip versus the spray tail is
attributed to premixed combustion that starts at the tip.
On the other hand, the spray tail experiences only
diffusion combustion. The air zone temperature profile
essentially reflects the compression process, reaching
its peak close to TDC, and dropping as the piston moves
down. The global cylinder temperature falls between the
temperatures of burned zones and the air zone, and is
considerably lower than the temperatures of burning or
burned zones. This explains why zero-dimensional or
single zone models are not suitable for emission
predictions.
Figure 12 shows predictions of the radiative, convective,
and total heat transfer rates over the duration of the
expansion process. It illustrates the importance of
radiant heat transfer in a diesel engine relative to the
convective and total heat transfer rates. The radiant heat
transfer rate increases rapidly right after the start of
ignition and the ratio of radiant to total heat transfer
remains higher than 30% throughout the process.

Figure 9. Predicted mass of air entrained into spray


zones, mass of air burned in spray zones and air
entrainment rate into spray zones at 1800 rpm, 50%
load.

Figure 10. Predicted mass fractions of fuel injected,


vaporized and burned versus crank angle at 1800
rpm, 50% load.

Model Validation Over a Range of Engine Loads


To verify the accuracy of the multi-zone model over a
wide range of operating conditions, engine performance
was simulated and compared with experimental data for
a range of loads varying from 25% to 100%, at a speed
of 1800 rpm. The values of the constants, B1 and B2 in
equation (38) and equation (39), previously determined
by the calibration process were kept constant throughout
the load range. Figures 7 and 13 compare predicted and
measured cylinder pressure traces at different loads.
Figures 8 and 14 compare apparent heat release rates
calculated from the predicted and measured pressure
traces. Both pressure and heat release rate comparisons
demonstrate that the multi-zone model is capable of
predicting the in-cylinder phenomena over a wide range
of engine loads, without changes in the values of the
calibration constants.
The predicted cylinder pressure traces have been
integrated to yield the indicated mean effective pressure
(IMEP) over the engine cycle. Using the Millington and
Hartles [60] friction model embedded into the cycle
simulation, the IMEP predictions have been converted
into Brake Mean Effective Pressures (BMEP).
Comparison of the predicted BMEP levels with
measured data at various load conditions indicates that
the model captures measurements with high fidelity,
except at the highest load condition (see Fig. 15). The
discrepancies in the latter condition are attributed to a
combination of diffusion heat release and friction model

Figure 11. Predicted temperatures of spray tip, tail,


and air zone compared with global gas temperature
at 1800 rpm, 50% load.

Figure 12. Simulation predictions of the radiative,


convective, and total heat transfer rates over the
duration of the expansion process

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

shortcomings, as well as difficulties in capturing the fuel


injection dynamics. In Figure 16, calculated and
measured ignition delays at each load condition are
plotted. The start of combustion observed in the
experiments was determined by the first peak of the
second derivative of the measured cylinder pressure, as
recommended by Assanis et al. [58]. Measured ignition
delays show excellent agreement with predicted values.
As load increases, exhaust gas temperature increases
with the richer engine operation. With more energy to be
recovered in the turbocharger, the resulting boost
pressure increases. Accordingly, cylinder pressure and
temperature during the ignition delay period increases.
Higher pressure and temperature with higher load

eventually decrease the ignition delay period.


Consequently, as load increases, the relative importance
of the premixed combustion phase decreases, with
combustion becoming more and more mixing-controlled,
as can be observed in Figures 8 and 14.

(a) 100% load

100% load

(b) 75% load

75% load

(c) 25% load

(c) 25% load

Figure 13. Comparison of predicted and measured


cylinder pressure traces over a range of loads at 1800
rpm.

Figure 14. Comparison of apparent heat release rates


computed from predicted and measured pressure
traces over a range of loads at 1800 rpm. Also shown
are fuel injection rates.

Model Validation Over a Range of Engine Speeds


The fidelity of the multi-zone model has been explored
further by comparing its predictions with data acquired
on the heavy -duty engine over a range of speeds at
constant load. Engine speed has been varied from 900
to 2100 rpm in increments of 300 rpm. Load was set at

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

50%. The values of the heat release constants, B1 and


B2 in equation (38) and equation (39) were frozen to their
initial calibration settings. Figures 7 and 17 compare
predicted and measured cylinder pressure traces at
different engine speeds. Figures 8 and 18 compare
corresponding apparent heat release rates calculated
from predicted and measured pressure traces. Overall,
both pressure and heat release rate comparisons show
that predictions over a wide range of engine speeds are
in good agreement with experimental data. Very
reasonable agreement, in terms of both magnitude and
trend, is also exhibited when comparing predicted and
measured ignition delays at various speeds in Figure 19.
Note that the heat release profile at 900 rpm exhibits a
more distinctive premixed combustion phase than the
other cases. This is attributed to the long ignition delay
observed for this case (see Fig. 19), which is primarily
caused by (i) the low injection rates and low injection line
pressures used, thus prolonging the physical preparation
of the fuel-air mixture and (ii) more time per cycle
available for heat loss, thus prolonging the chemical
aspects of the ignition delay. Since probably all fuel
mass is injected during the ignition delay period, this
results in an almost absent mixing-controlled, diffusion
phase heat release. A comparison of measured and
predicted BMEP levels in Figure 20 indicates that the
multi-zone model slightly underestimates BMEP. This is
primarily attributed to slightly lower predictions of
cylinder pressure than experimental data as well as
friction model shortcomings.

(a) 2100 rpm

(b) 1500 rpm

Figure 15. Comparison of predicted


measured data at various load conditions.

BMEP

with
(c) 1200 rpm

(d) 900 rpm


Figure 16. Predicted and measured ignition delays at
various load conditions.

Figure 17. Comparison of predicted and measured


cylinder pressure traces over a range of speeds at 50%
load.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

(a) 2100 rpm

(b) 1500 rpm

Figure 19. Predicted and measured ignition delays at


various engine speeds.

Figure 20. Comparison of predicted


measured data at various engine speeds.

BMEP

with

VALIDATION AGAINST SINGLE-CYLINDER


ENGINE EXPERIMENTS

(c) 1200 rpm

(d) 900 rpm


Figure 18. Comparison of apparent heat release rates
from predicted and measured pressure over a range of
speeds at 50% load. Also shown are fuel injection rates.

Additional validation of the multi-zone model's


predictions was conducted using reported data from an
independent engine set-up at the Engine Research
Center of the University of Wisconsin [61]. This work on
a modern single-cylinder, four-diesel stroke engine was
selected for the completeness in reporting engine
performance, soot and NOx emissions over a range of
injection timings. The engine set-up is based on a
version of the Caterpillar 3406 production engine with
simulated turbocharging, and is capable of producing 54
kW at a rated speed of 2100 rpm. Engine specifications
pertinent to this study are listed in Table 5.

Table 5. Engine specifications


Displacement volume (liter)
Bore (mm)
Stroke (mm)
Con. Rod length (mm)
Compression ratio (-)
Rated speed/power (rpm/kW)
Intake valve timing (degree CA)
Exhaust valve timing (degree CA)
Nozzle hole diameter (mm)
Number of nozzle holes (-)

2.44
137.19
165.1
261.62
15 : 1
2100/54
Open 3 ATDC
Close 10 ABDC
Open 19 BBDC
Close 7 BTDC
0.259
6

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

Test conditions are listed in Table 6. The operating


conditions were chosen to reflect common loads placed
on the engine (1600 rpm and equivalence ratio of 0.45,
or approximately 80% load), as well as the appropriate
inlet temperature and pressure conditions. Beyond the
baseline injection timing (11 deg. BTDC), predictions of
performance and emissions are compared with
experimental data for injection timings of 5, 8, 13 and 15
deg. BTDC. An additional objective is to explore the
sensitivity of the model to discretization parameters,
such as the zonal and time step resolution, so as to
optimize its numerical performance and robustness.
Table 6. Baseline operating conditions
Parameter
Value
Engine speed (rpm)
1600
Equivalence ratio
0.45
Injection timing (deg. BTDC)
11
Inlet air temperature (deg. C)
36
Inlet air pressure (kPa)
184
Exhaust back pressure (kPa)
159

(a)

Numerical sensitivity of multi-zone model


The sensitivity of model predictions to the selected zonal
and time step resolution was explored first so as to
optimize model robustness and numerical behavior. To
investigate the behavior of the multi-zone model with
respect to the zonal resolution, the baseline operating
point was simulated for different numbers of zones in the
radial direction. As can be seen in Figure 21a,
computational time increases almost linearly as the
number of zones is increased. Predicted IMEP data
show complete convergence to 1.3 MPa when 5 radial
zones are used. Note however, that even without any
resolution in the radial direction (one zone used), the
predicted IMEP is only 1% less. It can be concluded that
engine performance can be predicted with a minimal
number of zones in the radial direction without losing any
significant accuracy. On the other hand, predicted NO
and soot predictions are increased slightly (5.4% and
4.7%, respectively) as the number of zones is increased
from 1 to 5 (see Fig. 21b). Thereafter, NO and soot
values tend to converge, with differences being less than
2% compared to using a larger number of zones. As
expected, predicted emission values are more sensitive
to the number of zones used than IMEP. Therefore, it is
recommended to use at least 5 zones in the radial
direction for accurate prediction of NO and soot
emissions with acceptable computational time.
Another investigation on the behavior of the multi-zone
model with respect to the selected time step size was
conducted for the baseline operating condition. As can
be seen in Figure 22a, computational time increases
exponentially as time step is decreased. Predicted
IMEP, NO, and soot values (Figure 22b) are stable and
almost constant as time step is changed from 1 degree
CA to 0.5 degree CA. However, predicted values of the
parameters start diverging as time step is decreased
further to 0.25 degrees CA. This behavior is attributed to
truncation errors that start building-up as the time step
decreases because certain computed variables become
too small. Therefore, a time step of either 1 or 0.5
degrees CA is recommended for stability and
robustness. In this study, a time step size of 1 degree
CA was used for all predictions.

(b)
Figure 21. Model behavior with respect to the number
of zones used in the radial direction: (a) Normalized
computational time and IMEP, (b) Predictions of NO
and soot emissions.

(a)

(b)
Figure 22. Model behavior with respect to the time step
size used: (a) Normalized computational time and
IMEP, (b) Predictions of NO and soot emissions.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

Experimental
Conditions

Validation

under

Baseline

Operating

All model constants, including the constants of the twophase heat release model, B1 and B2 in equation (38)
and equation (39) have been kept at the same values as
previously determined through calibration against the
multi-cylinder diesel engine data. Figure 23 compares
the measured cylinder pressure trace with the one
predicted by the multi-zone model for the baseline
engine operating condition.
The measured peak
pressure is 9.64 MPa at 372.5 degrees CA and the peak
pressure predicted by the model is 9.61 MPa at 371
degrees CA. Overall it shows that the model predicts
pressure with good accuracy. In Figure 23 the predicted
apparent heat release rate is also plotted versus crank
angle. Since the method of calculating heat release rate
in Nehmer and Reitzs work is not specified, direct
comparison of heat release rate from experiment with
numerical result was not possible. The first peak of the
heat release rate is predicted to occur at 358 degrees
CA, while the second one occurs at 374 degrees CA.
Correspondingly, experimental data indicate that the two
peaks occur at 359 and 372 degrees CA, respectively. In
summary, the comparisons of cylinder pressures and
heat release rates under baseline operating conditions
indicate that the multi-zone spray combustion model
performs well for a second engine without change of the
calibration constants determined for the first engine.
Recall that the Extended Zeldovich mechanism has
been employed for NO predictions using standard rate
constants, as reported in Heywood [44]. Nitric oxide
forms throughout the high-temperature burned gases
behind the flame through chemical reactions involving
nitrogen and oxygen atoms and molecules, which do not
attain chemical equilibrium. The higher the burned gas
temperature is, the higher the rate of formation of NO is.
The critical time period is when burned gas temperatures
are at a maximum: i.e., between the start of combustion
and shortly after the occurrence of peak cylinder
pressure. Mixture which burns early in the combustion
process is especially important since it is compressed to
a higher temperature, increasing the NO formation rate,
as combustion proceeds and cylinder pressure
increases. After the time of peak pressure, burned gas
temperature decreases as the cylinder gases expand.
The decreasing temperature due to expansion and due
to mixing of high-temperature gas with air or cooler
burned gas freezes the NO chemistry. This second
effect means that freezing occurs more rapidly in the
diesel than in the spark-ignition engine, and much less
decomposition of the NO occurs. Figure 24 illustrates the
in-cylinder NO formation history predicted by the multizone model, and compares its cumulative value against
the measured engine-out NO emissions. The predicted
NO mass at the end of the expansion stroke is 4.09 mg,
which compares very favorably with the measured mass
of 3.90 mg. Since crank-angle resolved measurements
of pollutant concentrations are generally not available,
the comparison of the predicted NO profile throughout
the cycle with measurements was not possible.
Model predictions of soot emissions are based on either
the global rate equations for soot formation and
oxidation proposed by Hiroyasu et al. [19] or the NSC
soot oxidation model. For the Hiroyasu et al model, the
pre-exponential
constant, Af , in equation (55) is set here
-1
at 150 sec , which is, the same value used by Patterson
et al. [8]. The other pre-exponential constant, Ao, in
equation (56) is calibrated (A o=2250) to match the soot

level with measurements under the baseline operating


condition. On the other hand, the diameter of the soot
-9
particles for the NSC model is set to 4.5x10 m. In
Figure 25, the in-cylinder mass of soot predicted by the
global rate equation by Hiroyasu et al. [19] is compared
with the mass predicted by the NSC soot oxidation
model. Soot forms in the rich unburned-fuel-containing
core of the fuel sprays, within the flame region, where
the fuel vapor is heated by mixing with hot burned
gases. Soot then oxidizes in the flame zone when it
contacts
unburned
oxygen.
Consequently,
the
concentration of soot starts building up earlier than NO
during the early stages of combustion when rich
unburned mixture is present. This can be observed by
comparing the timings for the start of NO and soot
formation from Figure 24 and 25. After reaching its peak,
soot concentration decreases as oxygen is supplied by
air entrainment.
Soot destruction freezes later

Figure 23. Comparison of predicted and measured


cylinder pressure traces and predicted apparent heat
release rate. Injection timing is 11 deg. CA BTDC.

Figure 24. In-cylinder NO formation history predicted


by multi-zone model and measured exhaust NO mass
(Injection timing: 11 deg. CA BTDC).

Figure 25. In-cylinder soot formation histories predicted


by global rate equation vs. NSC model and measured
exhaust soot mass (Injection timing: 11 deg. CA
BTDC).

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

compared to NO dynamics since air entrainment


promotes soot oxidation. In general, the soot formation
and destruction trends predicted by the Hiroyasu et al.
and NSC models are quite similar. However, the NSC
model predicts a lower peak than the global rate model,
at least for this engine condition. Note that the measured
engine-out soot mass is 0.109 mg, while the predicted
soot mass by the global rate model is 0.106 mg versus
0.118 mg predicted by the NSC model.
Experimental Validation under Various Injection Timings
Following model validation under the baseline operating
condition, more extensive comparisons of predictions
with measurements were carried out over a range of
injection timings (from 5 to 15 degrees CA BTDC). The
values of all model constants remain unchanged. As
injection timing is advanced, cylinder pressure and
temperature during the ignition delay period become
lower. Therefore the ignition delay period becomes
longer. These phenomena result in more fuel burned
during the premixed combustion phase following the
ignition delay period, as can be observed in Figures 23
and 26. With combustion rates during the premixed
phase being much higher than those during the diffusion
phase, cylinder pressure increases faster as timing is
advanced. This effect is compounded through the
compression of cylinder gas by the rising piston, thus
increasing further cylinder pressure and temperature
with injection timing advances, as shown in Figure 27.
Figures 23 and 26 also indicate that predictions of
cylinder pressure traces are in very good agreement with
measurements over the entire range of injection timings
studied. The rising pressure levels with advancing
injection timings result in higher IMEP levels.
Consequently, predicted BMEP levels that are shown in
Figure 28 reflect this trend, with their magnitude being in
satisfactory agreement with measurements; errors are
attributed to friction model and injection profile
discrepancies.
Overall,
our
findings
are
very
encouraging given that the values employed for the
combustion model constants are the same as those
used for another similar engine.
Higher cylinder temperatures with advancing injection
timing generate more NO and less soot, resulting in the
perennial NO-soot tradeoff.
In Figure 29, NO-soot
tradeoff curves are compared with the two alternative
soot oxidation models and compared against
measurements over a range of injection timings from 5
to 15 degrees CA BTDC. While the general experimental
trend can be predicted by the multi-zone model using
either soot models, it appears that the Hiroyasu et al.
model predicts soot emissions better than the NSC
model across the range. In Figure 30, the measured and
calculated NO and soot emissions (by global rate
equations 55 and 56) are plotted individually as a
function of injection timing. The difficulty in predicting
more accurately measured NO emissions for this engine
appears to be responsible for the discrepancies in the
NO-soot trend observed in Figure 29. In contrast, the
global rate soot model emissions are in excellent
agreement with measurements.

CONCLUSIONS
A quasi-dimensional, multi-zone, direct injection diesel
spray combustion model has been developed and

implemented in a full cycle simulation of a turbocharged


engine for the purpose of predicting engine performance
and NO and soot emissions.
A comprehensive
validation of the proposed multi-zone model against
experiments in a constant volume chamber, as well as
single and multi-cylinder engines has been conducted
over a range of injection pressures, injection timings,
engine speeds and engine loads. The major conclusions
are the following:

A computational time step of either 1 or 0.5


degrees CA is recommended for numerical
stability and robustness. In this study, a time
step size of 1 degree CA was used for all
predictions.

Engine performance can be predicted accurately


with a minimal number of zones in the radial
direction.
However, as expected, predicted
emissions are more sensitive to the number of
zones used than IMEP predictions. At least 5
zones in radial direction are recommended for
prediction of NO and soot emissions.

A modified Hiroyasu correlation for spray tip


penetration correlation shows good agreement
with experimental data over a range of injection
pressures in a constant volume vessel. While
the Reitz and Bracco correlation for the spray
angle produces predictions in acceptable
agreement with data, an improved correlation is
needed to describe more accurately the spray
shape.

The ignition delay, premixed and diffusion


combustion models
have been shown to
predict rates of heat release and engine
performance
with
high
fidelity,
over
a
comprehensive range of engine loads, speeds
and injection timings for two different engines.
Calibration of two critical model constants was
limited to only one operating condition, a finding
that is very encouraging.

The extended Zeldovich scheme and the global


rate, two-equation model of Hiroyasu et al.
appears to predict the correct trends for the NOsoot trade-off for at least one engine. Further
investigation is required to enhance the fidelity
of NO and soot emissions predictions across a
wider range of engine and operating conditions.

ACKNOWLEGEMENTS
The authors would like to acknowledge the technical and
financial support of the Automotive Research Center
(ARC) by the National Automotive Center (NAC) located
within the US Army Tank-Automotive Research,
Development and Engineering Center (TARDEC). The
ARC is a U.S. Army Center of Excellence for Automotive
Research at the University of Michigan, currently in
partnership with University of Alaska-Fairbanks,
Clemson University, University of Iowa, Oakland
University, University of Tennessee, Wayne State
University, and University of Wisconsin-Madison.
Timothy J. Jacobs, Pin Zeng and Zoran Filipi are
gratefully acknowledged for providing experimental data
acquired on the ARC heavy -duty engine test cell for use
in model validation.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

(a) Injection timing: 5 deg. CA BTDC

Figure 27. Comparison of predicted cylinder global


temperature profiles over a range of injection timings.

(b) Injection timing: 8 deg. CA BTDC


Figure 28. Comparison of predicted and measured
BMEP over a range of injection timings.

(c) Injection timing: 13 deg. CA BTDC


Figure 29. Comparison of predicted and measured NOsoot tradeoff curves for injection timings ranging from 5
to 15 deg. CA BTDC.

(d) Injection timing: 15 deg. CA BTDC


Figure 26. Comparison of predicted and measured
cylinder pressure traces. Also shown are predictions of
the apparent heat release rate.

Figure 30. Comparison of predicted and measured NO


and soot emissions over a range of injection timings.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

The expression for can be obtained from the condition


of S=lb at t=tb.

APPENDIX

=2

MODIFIED HIROYASUS CORRELATION FOR


SPRAY PENETRATION
Levich [62] derived the breakup length of a liquid jet at
high velocities under the following assumptions:

The liquid jet of density, l, is moving in a gas


medium with a density, a.
The relative velocity between the liquid jet and
the gas medium is large.

The amplitude of the jet surface disturbance is


accelerated by the pressure disturbance of the
gas.

As the amplitude of the jet increases, the jet


tends to become unstable, and finally, the jet
may break up into droplets.

The breakup length, lb, is given by:

lb u b tb

l
d
a n

(A1)

where ub is jet velocity within the intact core; tb is


breakup time; is a constant; dn is nozzle hole diameter.
If the pressure upstream of the injector nozzle can be
estimated or measured, and assuming the flow through
each nozzle is quasi steady, incompressible, and one
dimensional, the mass flow rate of fuel injected through
& i is given by [44]:
the nozzle m

m& i = CD An 2 l P

(A2)

& i = l An ui , the fuel injection velocity ui can be


Since m
expressed as:
2P

u i = C D
l

0. 5

(A3)

With the assumption that the jet velocity within the intact
core, ub, is equal to the initial jet velocity, ui, the breakup
time in equation (A1) can be expressed as follows:

tb =

l d n
2C D 2 a P

(A4)

Assuming that the value of the discharge coefficient can


be determined, the spray penetration before breakup is:

0 .25

(CD d n )
a
0. 5

0 .25

(A7)

Hiroyasu and Arai [45] determined the constants from


their experimental data as =15.8 and CD is 0.39. So,
the spray tip penetration after breakup becomes:

S = 2.95

0 .25

( d n t ) 0 .5

(A8)

Schihl et al. [33] developed a phenomenological cone


penetration model and compared their model with
Hiroyasus model and several experimental data sets
from different sources. Discharge coefficients of the
nozzles in the experimental data sources ranged from
0.64 to 0.74. Since the model of Schihl et al. is not valid
before or near the break-up region, they compared the
penetration after the breakup time. The comparison
indicated that the model of Hiroyasu and Arai shows
fairly good agreement with experimental data sufficiently
after the breakup time.
Dan et al. [34] investigated spray characteristics as a
function of various injection pressures and ambient
conditions. They measured temporal change in spray tip
penetration with injection pressure, and compared their
data with empirical correlations proposed by Hiroyasu
and Arai [45], Wakuri et al. [63], and Dent [64]. The
measured discharge coefficient of the nozzle used by
Dan et al. is 0.66. Discharge coefficient is a parameter
that can be varied in the correlation of Wakuri et al.
However, discharge coefficient is not a variable in the
other two correlations. Instead, discharge coefficient
values of 0.8 and 0.39 are already embedded in the
correlations of Dent and Hiroyasu and Arai, respectively.
Comparison of measured spray penetration with
predicted data by three correlations shows that the
correlation proposed by Hiroyasu and Arai [45] fits better
than the other correlations with the experimental data of
Dan et al. [34]. From the above literature review, it is
concluded that the correlation for spray penetration after
breakup, equation (A8), is still valid even for nozzles with
different
discharge
coefficients.
However,
the
penetration correlation before breakup, equation (A5),
cannot be used with fixed discharge coefficient 0.39,
since the intact core velocity, which is assumed to be
equal to the injection velocity, is directly a function of
discharge coefficient. If the discharge coefficient is
different from 0.39, injection velocity and injection
pressure cannot be matched correctly. So, a modified
penetration correlation is proposed in this research.
Assuming that the value of the discharge coefficient is
predetermined, spray penetration before breakup should
be given by equation (A5).

0 .5

2P
t
S = C D

(A5)

Based on continuous jet theory, the spray tip penetration


1/2
is assumed to be proportional to t after breakup. The
spray tip penetration, S, can be expressed as follows:

S= t

(A6)

Since equation (A8) is supposed to be valid even for


discharge coefficient values other than 0.39, can be
expressed in terms of CD from equations (A6), (A7), and
(A8) as:

2.95 2
2 0 .5 C D

(A9)

By inserting this equation into equation (A4), the breakup


time, tb, can be obtained as follows:

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

t b = 4.351

l d n
C D ( a P) 0. 5
2

(A10)

Finally, the spray penetration from a nozzle with known


discharge coefficient, CD , can be described by equation
(A5) before breakup, and equation (A8) after breakup,
with the breakup time specified by equation (A10).

19.

20.

REFERENCES
1.

2.
3.
4.
5.
6.

7.

8.

9.
10.
11.
12.

13.
14.
15.
16.
17.
18.

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 Studies, SAE Paper 860329, 1986.
Krieger, R. B. and Borman, G. L., The Computation of
Apparent Heat Release from Internal Combustion
Engines, ASME Paper 66-WA/DGP-4, 1966.
Foster, D. E., An Overview of Zero-Dimensional
Thermodynamic Models for I.C. Engine Data Analysis,
SAE Paper 852070, 1985.
Oran, E.S. and Boris, J. P., Detailed Modeling of
Combustion Systems, Prog. Energy Combust. Sci., Vol.
7, pp. 1-72, 1981.
Bracco, F. V., Modeling of Engine Sprays, SAE Paper
850394, 1985.
Amsden, A. A., Ramshaw, J. D., ORourke, P. J., and
Dukowicz, J. K., KIVA: A Computer Program for Two- and
Three-Dimensional Fluid Flow with Chemical Reactions
and Fuel Sprays, Los Alamos National Laboratory Report
LA-10245-MS, 1985.
Amsden, A. A., Butler, T. D., and ORourke, P. J.,KIVA-II
Computer Program for Transient Multidimensional
Chemically Reactive Flows with Sprays, SAE Paper
872072, 1987.
Patterson, M. A., Kong, S. C., Hampson, G. J., and Reitz,
R. D., Modeling the Effects of Fuel Injection
Characteristics on Diesel Engine Soot and NOx
Emissions, SAE Paper 940523, 1994.
Varnavas, C. and D. N. Assanis, A High Temperature and
High Pressure Evaporation Model for the KIVA-3 Code,
SAE Paper 960629, 1996.
Austen, A. E. W. and Lyn, W. T., Some Steps Toward
Calculating Diesel Engine Behavior, SAE Paper 409A,
1961.
Whitehouse, N. D. and Sareen B. K., Prediction of Heat
Release in a Quiescent Chamber Diesel Engine Allowing
for Fuel/Air Mixing, SAE Paper 740084, 1974.
Shahed, S. M., Chiu, W. S. and Lyn, W. T., A
Mathematical Model of Diesel Combustion, Proceedings
of the Institution of Mechanical Engineers, C94/75, pp.
119-128, 1975.
Chiu, W. S., Shahed, S. M. and Lyn, W. T., A Transient
Spray Mixing Model for Diesel Combustion, SAE Paper
760128, 1976.
Hiroyasu, H. and Kadota, T., Models for Combustion and
Formation of Nitric Oxide and Soot in Direct Injection
Diesel Engines, SAE Paper 760129, 1976.
Merguerdichian, M. and Watson, N., "Prediction of Mixture
Formation and Heat Release in Diesel Engines," SAE
Paper 780225, 1978.
Dent, J. C. and Mehta, P.S., "Phenomenological
Combustion Model for a Quiescent Chamber Diesel
Engine," SAE Paper 811235, 1981.
Kamimoto, T., Kobayashi, H. and Matsuoka, S., A Big
Size Rapid Compression Machine for Fundamental
Studies of Diesel Combustion, SAE Paper 811004, 1981.
Kobayashi, H., Kamimoto, T. and Matsuoka, S.,
Photographic and Thermodynamic Study of Diesel

21.

22.
23.

24.
25.
26.
27.

28.

29.

30.

31.

32.

33.

34.
35.
36.

Com bustion in a Rapid Compression Machine, SAE


Paper 810259, 1981.
Hiroyasu, H., Kadota, T. and Arai, M., Development and
Use of a Spray Combustion Modeling to Predict Diesel
Engine Efficiency and Pollutant Emissions (Part 1
Combustion Modeling), Bulletin of the JSME, Vol. 26, No.
214, pp. 569-575, 1983a.
Hiroyasu, H., Kadota, T. and Arai, M., Development and
Use of a Spray Combustion Modeling to Predict Diesel
Engine Efficiency and Pollutant Emissions (Part 2
Computational Procedure and Parametric Study), Bulletin
of the JSME, Vol. 26, No. 214, pp. 576-583, 1983b.
Kono, S., Nagao, A. and Motoka H., Prediction of InCylinder Flow and Spray Formation Effects on
Combustion in Direct Injection Diesel Engines, SAE
Paper 850108, 1985.
Kyriakides, S. C., Dent, J. C., and Mehta, P. S.,
Phenomenological Diesel Combustion Model Including
Smoke and NO Emission, SAE Paper 860330, 1986.
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.
Bazari, Z., A DI Diesel Combustion and Emission
Predictive Capability for Use in Cycle Simulation, SAE
Paper 920462, 1992.
Li, Q. and Assanis, D. N., A Quasi-Dimensional
Combustion Model for Diesel Engine Simulation, ASME,
ICE-Vol. 20, 1993.
Yoshizaki, T., Nishida, K. and Hiroyasu, H., Approach to
Low NOx and Smoke Emission Engines by Using
Phenomenological Simulation, SAE Paper 930612, 1993.
Morel, T. and Wahiduzzaman, S., "Modeling of Diesel
Combustion and Emissions," 96 FISITA Proceedings, 26th
International Congress, Praha, Czech Republic, June 17 21, 1996.
Kouremenos, D. A., Rakopoulos, C. D., and Hountalas, D.
T., Multi-Zone Combustion Modeling for the Prediction of
Pollutants Emissions and Performance of DI Diesel
Engines, SAE Paper 970635, 1997.
Bhaskar, T. and Mehta, P. S., A Multi-Zone Diesel
Combustion Model Using Eddy Dissipation Concept,
Proceedings of the Fourth International Symposium on
Diagnostics and Modeling of Combustion in IC engines,
COMODIA 98, Kyoto, Japan, 1998.
Rakopoulos C. D. and Hountalas, D. T., Development
and Validation of a 3-D Multi-Zone Combustion Model for
the Prediction of DI Diesel Engines Performance and
Pollutants Emissions, SAE Paper 981021, 1998.
Gao, Z. and Schreiber, W., A Multizone Analysis of Soot
and NOx Emission in a D.I. Diesel Engine as a Function of
Engine Load, Wall Temperature, and Intake Air O2
Content, ASME Paper, 2000-ICE-314, 2000.
Tauzia, X., Hetet J.-F., Chesse, P., Inozu, B., and Roy, P.,
The Use of Phenomenological, Multizone Combustion
Model to Investigate Emissions from Marine Diesel
Engines, ASME Paper 2000-ICE-325, 2000.
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.
Dan, T., Takagishi, S., Senda, J., and Fujimoto, H., Effect
of Ambient Gas Properties for Characteristics of NonReacting Diesel Fuel Spray, SAE Paper 970352, 1997.
Overbye, V. D., Bennethum, J. E., Uyehara, O. A., and
Myers, P. S., Unsteady Heat Transfer in Engines, SAE
Trans. Vol. 69, 1961.
Annand, J. D., Heat Transfer in the Cylinders of
Reciprocating Internal Combustion Engines, Proc. Inst.
Mech. Eng., Vol. 177, No. 36, 1963.

Downloaded from SAE International by University of New South Wales, Friday, August 26, 2016

37. Ebersole, G. D., Myers, P. S., and Uyehara, O. A.,


38.

39.

40.
41.
42.
43.
44.
45.
46.

47.
48.
49.
50.
51.
52.

Radiant and Convective Components of Diesel Engine


Heat Transfer, SAE Paper 701C, 1963.
Annand, J. D., and Ma, T. H., Instantaneous Heat
Transfer Rates to the Cylinder Head Surface of a Small
Compression Ignition Engine, Proc. Inst. Mech. Eng., Vol.
185, No. 72, 1971.
Flynn, P., Mizusawa, M., Uyehara, O. A., and Myers, P.
S., An Experimental Determination of the Instantaneous
Potential Radiant Heat Transfer within an Operating Diesel
Engine, SAE Paper 720022, 1972.
Oguri, T. and Shigewo, I., Radient Heat Transfer in Diesel
Engines, SAE Paper 720023, SAE Trans., Vol. 81, 1972.
Sitkei, G. and Ramanaiah, G. V., A Rational Approach for
Calculation of Heat Transfer in Diesel Engines, SAE
Paper 720027, 1972.
Kunitomo, T., Matsuoka, K., and Oguri, T., Prediction of
Radiative Heat Flux in a Diesel Engine, SAE Paper
750786, SAE Trans., Vol. 84, 1975.
Dent. J. C. and Suliaman, S. J., Convective and Radiative
Heat Transfer in a High Swirl Direct Injection Diesel
Engine, SAE Paper 770407, 1977.
Heywood, J. B., Internal Combustion Engine
Fundamentals, McGraw-Hill Book Co., 1988.
Hiroyasu, H., and Arai, M., Fuel Spray Penetration and
Spray Angle of Diesel Engines, Trans. of JSAE, Vol. 21,
pp. 5-11, 1980.
Reitz, R. D. and Bracco, F. B., On the Dependence of
Spray Angle and Other Spray Parameters on Nozzle
Design and Operating Conditions, SAE Paper 790494,
1979.
Hiroyasu, H., Arai, M., and Tabata, M., Empirical
Equations for the Sauter Mean Diameter of a Diesel
Spray, SAE Paper 890464, 1989.
Borman, G. L. and Johnson, J. H., Unsteady Vaporization
Histories and Trajectories of Fuel Drops injected into
Swirling Air, SAE Paper 598C, 1962.
Gosman, A. D. and Johns, R. J. R., Computer Analysis of
Fuel-Air Mixing in Direct Injection Engines, SAE Paper
800091, 1980.
Ranz, W. E. and Marshall, W. R., Evaporation from
Drops, Chem. Eng. Prog., Vol. 48, No. 4, pp. 173-180,
1952.
Watson, N., Pilley, A. D. and Marzouk, M. A., Combustion
Correlation for Diesel Engine Simulation, SAE Paper
800029, 1980.
Nishida, K. and Hiroyasu, H., Simplified ThreeDimensional Modeling of Mixture Formation and
Combustion in a D.I. Diesel Engine, SAE Paper
890269,1989.

53. Lavoie, G. A., Heywood, J. B., and Keck, J. C.,

54.
55.

56.
57.

58.

59.

60.
61.

62.
63.

64.

Experimental and Theoretical Investigation of Nitric Oxide


Formation in Internal Combustion Engines, Combust. Sci.
Technol., Vol. 1, pp. 313-326, 1970.
Nagle, J. and Strickland-Constable, R. F., Oxidation of
Carbon between 1000-2000 C, Fifth Carbon Conference,
Pergamon, Oxford, Vol. 1, pp. 154-164, 1962.
Nishimura, A. and Assanis, D. N., A Model for Primary
Diesel Fuel Atomization Based on Cavitation Bubble
Collapse Energy, Eighth International Conference on
Liquid Atomization and Spray System (ICLASS), 2000.
Dan, T., The Turbulent Mechanism and Structure of
Diesel Spray, Ph. D. Thesis, Toshisya University, 1996.
Filipi, Z. S., Homsy, S. C., Morrison, K. M., Hoffman, S. J.,
Dowling, D. R., and Assanis, D. N., Strain Gage Based
Instrumentation for In-Situ Diesel Fuel Injection System
Diagnostics, ASEE Annual Conference, Milwaukee,
Wisconsin, June 15-18, 1997.
Assanis, D. N., Filipi, Z. S., Fiveland, S. B., and Syrimis,
M., A Predictive Ignition Delay Correlation under SteadyState and Transient Operation of a Direct Injection Diesel
Engine, ASME paper, 99-ICE-231, 1999.
Fiveland, S. B., Development of Engine Measurement
Techniques with Application to Steady State and Transient
Ignition Delay and Heat Release Analysis in a DirectInjection Diesel Engine, Masters Thesis, The University
of Michigan, 1999.
Millington, B. W., Hartles, E. R., Frictional Losses in
Diesel Engines, SAE Paper 680590, SAE Transactions,
Vol. 77, 1968.
Nehmer, D. A. and Reitz, R. D., Measurement of the
Effect of Injection Rate and Split Injections on Diesel
Engine Soot and NOx Emissions, SAE Paper 940668,
1994.
Levich, V. G., Physicochemical Hydrodynamics, PrenticeHall Inc., Englewood cliffs, New Jersey, pp. 639-650,
1962.
Wakuri, Y., Fujii, M., Amitani, T., and Tsuneya, R.,
Studies of the Penetration of a Fuel Spray in a Diesel
Engine, Bulletin of the JSME, Vol. 3, No. 9, pp. 123-130,
1960.
Dent, J. C., Basis for the Comparison of Various
Experimental Methods for Studying Spray Penetration,
SAE Paper 710571, 1971.

You might also like