Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

SAE TECHNICAL

PAPER SERIES 980131

Modeling the Effects of Fuel Spray


Characteristics on Diesel Engine
Combustion and Emission
Mark A. Patterson and Rolf D. Reitz
University of Wisconsin-Madison

Reprinted From: Advances in Multi-Dimensional Modeling


(SP-1329)

International Congress and Exposition


Detroit, Michigan
February 23-26,1998

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760
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 Sec-
tions 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 1998 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
980131

Modeling the Effects of Fuel Spray Characteristics on Diesel


Engine Combustion and Emission
Mark A. Patterson and Rolf D. Reitz
University of Wisconsin-Madison

Copyright 1998 Society of Automotive Engineers, Inc.

ABSTRACT ensure a rigorous test of the new droplet breakup model.


The array of conditions chosen ensures that the model
A new spray model has been developed to improve the will function well under diesel-engine conditions.
prediction of diesel engine combustion and emissions
using the KIVA-Il CFD code. The accuracy of modeling Table 1. Specifications for the experimental study of
the spray breakup process has been improved by the Alloca et al. (1995)
inclusion of Rayleigh-Taylor accelerative instabilities,
Experiment Type Constant Volume Bomb
which are calculated simultaneously with a Kelvin-Helm-
Injector Type Single Hole Injector with PE-Bosch Injection
holtz wave model. This model improves the prediction of Pump
the droplet sizes within a diesel spray and provides a Chamber Gas Air
more accurate initial condition for the evaporation, com- Number of Holes 1
bustion, and emissions models. An improvement to the Hole Diameter 0.200 mm
droplet drag model is also presented. This model L/D ratio of holes 3.0
accounts for the increased droplet drag due to the Injection Pressure 98 MPa (max.)
change in the droplet's shape, as well as the increase in Chamber Pressure 1.70 MPa
the frontal area of the droplet. The drag model affects the
breakup process locally, producing a more realistic drop- EXPERIMENTAL VALIDATION DATA
let size distribution, and therefore a more accurate calcu-
lation of the vaporization process. Through the The experimental studies used for validation of the model
introduction of this model the prediction of the pressure, are comprised of constant-volume bomb and diesel
heat release, and emissions produced by single and split engine studies. The first constant-volume bomb case is
injections can now be accurately modeled. that of Alloca et al. (1992). This study provides experi-
mental data for the local droplet Sauter Mean Diameter
INTRODUCTION (SMD), which was measured for all locations within the
spray at a given instant for times between 0.2 and 1 milli-
Modeling diesel engines is rapidly becoming an impor- second. The measurements were accomplished using
tant avenue of research. Effective models which provide the light extinction technique, which is presented in detail
costefficient ways of studying different engine geome- in Su (1995). The data, with the conditions summarized
tries, operating conditions, and injection strategies are in Table 1, provide the most severe test of the spray
essential tools in modern engine design. A good working model because both the temporally and spatially
model can reduce the number of experimental test cases resolved size distributions within the spray are available
which must be performed to meet increasingly stringent for comparison to the KIVA predictions.
worldwide emissions standards. To this end, a new spray The second constant-volume bomb experiment consid-
model has been added to the KIVA-II code, and exten- ered the relationship between injection pressure and soot
sively tested to ensure that the prediction of the spray formation (Kamimoto et al. 1987). A rapid compression
produces the conditions needed for the correct prediction machine was used to study sprays in non-evaporating,
of evaporation, combustion, and emissions formation. evaporating, and combusting conditions by varying the
To verify the behavior of the new spray model, amount of oxygen and the temperature of the chamber.
experimental studies which include non-evaporating The focus, for the purposes of this study, is on the pene-
sprays, evaporating sprays without combustion, and die- tration of the liquid and gaseous phases for an evaporat-
sel engine cases were used. The local dropsize, liquid ing spray without combustion. Table 2 gives the
penetration, vapor penetration, pressure, heat release, specifications range of pressures were studied, but the
and emissions were studied in comparison to experi- case chosen for comparison was 80 MPa, since it is seen
ments. Each of these studies has been selected to as the minimum for future engine development. Both the

1
piston and the head of the chamber were transparent, so and 90 MPa for the single injection cases, and 9OMPa for
that the spray injected across the chamber could be the split injection cases.
seen, and the height of the pancake-shaped section was
40 mm. Both combusting and non-combusting sprays Table 3. Specifications for the engine used in Espey et
were studied by using air or nitrogen in the cylinder. al. (1994)
Engine Type Cummins Single-Cylinder,
Table 2. Specifications for the spray bomb used in Four-Stroke, DI Diesel
Kamimoto et al. (1987) Bore 140 mm
Stroke 152 mm
Engine Type Rapid Compression Machine
Displacement 2.34 liters
Bore 196 mm
Connecting Rod Length 305 mm
Stroke 560 mm
Compression Ratio 11:1 (Heated/Pressurized Intake Air)
Compression Ratio 14.7
Injector Type Closed-Nozzle, Unit Injector
Injector Type Hydraulically actuated Electronically
controlled Unit Injector (HEUI) Peak Injection Pressure 68 MPa
Number of holes 1 Number of Holes 8
Hole Diameter 0.16 mm Hole Diameter 0.203 mm
L/D ratio of holes 3.125 L/D of Holes 3.9
Fuel Type n-tridecane (C13H28) Angle of Fuel-Jet Axis 14 deg.
(from horiz.)
Injection Pressure 80 MPa

The measurements taken include spray cone angle, Table 4. Specifications for the Caterpillar engine used
spray tip penetration, spray volume, air excess ratio, Sau- in Nehmer (1993) and Tow et al. (1994).
ter mean diameter, and vapor and liquid penetration. The Bore 137.19 mm
vaporizing spray was injected into an environment at 3.1 Stroke 165.1 mm
MPa and 900 K which was composed of nitrogen to pre- Connecting Rod Length 261.62 mm
vent combustion. Displacement 2.44 L
Compression Ratio 15.0
The optical engine used Planar Laser-Induced Rayleigh
Piston Crown Mexican hat
scattering (PLIR) in a Cummins diesel engine and results Engine Speed 1600 RPM
of spray penetration, pressure, heat release, and some Intake pressure 184 kPa
images of the equivalence ratio are available (Espey et Intake temperature 310 K
al. 1994). The engine had a "square" bowl with transpar- Fuel Injected 0.1622 g/cycle
ent sides to allow the laser sheet to be directed along the Peak Injection Pressure 90 MPa
axis of one of the spray plumes in the engine. The bowl Intake valve close timing 147 deg. BTDC
was flat piston that had a flat-sided annulus of transpar- Swirl ratio (nominal) 1.0
ent material on top of it. Overall, the results of Nehmer's study show that the
The engine had an 11:1 compression ratio, with heated soot-NOx tradeoff becomes flattened with split injections,
intake air compressed to 206 kPa to make the conditions and particulate emissions are produced with values of
at TDC match the production engine's values of 1050 K approximately 0.07 to 0.09 g/hp-hr. The NOx emissions
and 5 MPa for the motored engine. The engine's specifi- can be reduced by changing the rate of injection and how
cations axe listed in Table 3. This engine had almost the much fuel is injected in the first pulse.
same penetration length of the liquid spray as that of
Kamimoto et al. Thus, in high-pressure engine environ- MODEL FORMULATION
ments the spray penetrates the same distance, regard-
less of engine geometry, injection pressure, or whether KELVIN- HELMHOLTZ MODEL THEORY The TAB
combustion takes place. model has been replaced with the wave spray breakup
model in the present version of KIVA. The wave model
Extensive studies of a Caterpillar engine have given
(Reitz and Bracco, 1986) assumes that breakup is
pressure, heat release, emissions, and other engine per-
caused by instabilities on the surface of a cylindrical 'blob'
formance data over a wide range of operating conditions
of liquid. In this analysis, the coordinate system is cylin-
(Nehmer, 1993), (Tow et al., 1994). In Nehmer's study,
drical and fixed with respect to the liquid. Thus, the case
the engine was run with single and split injections, and
of fuel injection is viewed as a high velocity gas moving
was shown to produce the most favorable emissions
over a stationary liquid blob. Both the liquid and the gas
when the fuel was injected in multiple pulses, with a sig-
are assumed to be incompressible, and the gas is
nificant dwell between pulses. The specifications for this
assumed to be inviscid.
engine are given in Table 4. A full dilution tunnel was
used for particulate measurements, and both dry soot
and SOF (using Soxhlet extraction) data were recorded
for this set of experiments. The injector is a common rail
system (Miyaki et al., 1991) with injection pressures of 45

2
The mechanical relations are combined with the lin- nozzle, and is allowed to have a radius larger than the
earized hydrodynamical equations to form a complete nozzle for the initial breakaway from the liquid core of the
description of the motion of the liquid and the gas. The spray. This is implemented as follows:
solution can be found by transforming the equations of
motion into stream and potential functions. The result is
the dispersion relationship
(Eq. 5)
where B0 is taken as 0.61.

RAYLEIGH-TAYLOR MODEL THEORY In the original


presentation of the Taylor (1963) model, the theory was
based on the idea that acceleration normal to the inter-
(Eq. 1)
face between two fluids of different densities can cause
where L2 = 2 + l and the I and K terms are Bessel instabilities under certain conditions. If the two fluids are
functions, is the wave growth rate, is the wave num- a liquid and a gas, for example, the interface between
ber, and is the liquid viscosity. them is stable when the acceleration is directed into the
liquid.
This function was fitted numerically to obtain the wave-
length and the frequency for the fastest growing wave on The model proposed by Chang (1991) is based on the
the surface (Reitz, 1987). The curve fits obtained by this original Taylor model with the assumption of linearized
analysis for the wavelength and the growth rate are disturbance growth rates and negligible viscosity (Bell-
man and Pennington, 1954). The problem of disturbance
growth is solved in two dimensions with the acceleration
of the fluids normal to the interface between them, and
(Eq. 2) both fluids are assumed to be incompressible with an infi-
nite interface length and infinite depth.
In the analysis, the acceleration of the droplet was found
(Eq. 3)
by dividing the drag force by the mass of the droplet. This
where We is the Weber number is for the gas, defined as allows a relationship between the drop acceleration and
with ur as the relative velocity between the drag coefficient, which is
the liquid and the gas, and as the surface tension of the
liquid. The Ohnesorge number, Z, is defined as
with the liquid Weber number, We, simi- (Eq. 6)
lar to We, except that the liquid density is used, and the
Taylor number, T, is defined as . where a is the droplet acceleration, ur is the relative
velocity between the droplet and the air, and r is the
The breakup time satisfies two conditions: bag breakup radius of the droplet.
at low Weber numbers, and stripping breakup at high
Weber numbers (Reitz and Diwakar, 1987). Both of these The equations for the wavelength and frequency of the
limiting cases are observed for a breakup time , which is fastest growing waves are:
expressed as

(Eq. 7)
(Eq. 4) and
with B1 as an arbitrary constant. This constant accounts
for differences in the drop breakup time due to unknown
initial conditions for the drop. The influence on B1 should (Eq. 8)
be related to upstream conditions in the nozzle, such as
The breakup time is found by taking the reciprocal of the
L/D ratio, turbulence inside the injector, and cavitation at
frequency of the fastest growing wave, as given in equa-
high injection pressures, for example. This constant has
tion (8). Substituting equation (6) for 'a' in equation (8)
been set as low as 1.73, based on comparisons to the
and simplifying allows the calculation of the dimension-
TAB model (O'Rourke and Amsden, 1987) and as high as
less time constant for breakup:
60 for some calculations.
The drop size of the newly formed droplet, r, is assumed
to be proportional, to the calculated wavelength, unless
the wavelength is large compared to the drop cir- (Eq. 9)
cumference. In the case of large wavelength, the droplet where the Weber number for the air is .
is assumed to be forming from the liquid column near the

3
This dimensionless time was used to calculate the of the spray, the overestimation of shed mass frequently
breakup timescale of each droplet. The droplet was only led to mass conservation errors. The amount of shed
allowed to break up if the wavelength given by equation mass could actually become higher than the mass of the
(7) is smaller than the droplet diameter. A simple and parent parcel for parcels with a small initial mass. For
physically realistic model is to allow a competition these reasons, the old method of summation was not
between the wave and Rayleigh-Taylor breakup models used in this study. Even when the model is listed as the
for each droplet, and to break the droplet up when one of 'previous' model, the calculation of shed mass from equa-
the three predicts breakup should occur. The implemen- tion 13 is used to ensure mass conservation.
tation of this model is given next.
When a new parcel is created, it is important to make
sure that the size distribution after the breakup event cor-
MODIFIED WAVE BREAKUP MODEL To improve the
responds to that given by equation (10) since that result
wave breakup model, it was of interest to make sure that
is based on experimental data. In previous implementa-
the droplet sizes were consistent with the original model
tions of the wave model, a new droplet was created with
of Reitz and Diwakar (1987). In that model, the size of the
the radius proportional to the wavelength, and the num-
droplet was calculated as follows:
ber of droplets was calculated based on the total mass
that had been shed from the original parcel. Using this
approach, the average radius (SMR) of the two parcels
(Eq. 10)
after breakup is less than what is predicted by equation
where r is the radius of the droplet, r is the change in the (10) when the droplets are regrouped from the parent
droplet size in one time step, and is given by equation parcel into the two child parcels.
(4).
To maintain the SMR consistent with equation (10) after
In the present breakup model, care was taken to ensure breakup, the new 'parent' parcel must actually contain
mass conservation was respected after breakup. A previ- droplets that are larger than those in the previous parent
ous model used by Su et al. (1995) accounted for the parcel. Physically, this can be justified if the radius of the
mass of small droplets shed from the parents using the droplets in a parcel is assumed to represent an average
equation: of the sizes within the parcel. Creating new parcels repre-
sents grouping the smaller droplets into the newly cre-
ated parcel, leaving the larger droplets in the parent
parcel. Thus, in this concept, the creation of new parcels
(Eq. 11) is not a direct analog to breakup, but a regrouping of the
where mshed is the mass shed from the parent parcel, droplets after the breakup event takes place.
is the density of the liquid fuel, ni is the number of drop- The breakup (or parcel division) process, thus, must obey
lets in the parent parcel at time step i, and r is the radius liquid mass conservation and conservation of SMR. The
of the droplets in the parent parcel at the subscripted mass balance is:
time step. The summation starts at i = 0 which is the time
where the droplet begins the breakup process, and is
continued up to term j, where enough shed mass exists (Eq. 14)
to create a new parcel. This is 3% of the mass of an
injected parcel in both the previous and present versions where is the fuel density, n is the number of droplets in
of the spray model. the parcel, r is the radius of the droplets in the parcel, and
the subscripts are p for parent droplet, b for the 'bigger'
The formulation of equation 11 produces an overesti- radius child droplet (i.e., the new parent droplet), and s
mation of the amount of shed mass. A simple expansion for the 'smaller' radius child droplet. Using the same nota-
of terms for j = 2 helps to show this: tion, the conservation of SMR is:

(Eq. 12)
(Eq. 15)
Since the number of droplets in the parent parcel remains
at n0 after removing the shed mass, the amount of the Substituting equation (14) into (15):
shed mass should be:
(Eq. 16)

(Eq. 13) The radius of the smaller child droplet is assumed to be


proportional to the size of the Kelvin-Helmholtz wave,
This is simply the change in mass of the parent droplets. obtained from equation (5). The number of bigger child
Upon closer inspection, noting that n2 > n1 > n0, it is clear (or new parent) droplets is equal to the number of parent
that equation (12) overestimates the shed mass. In addi- droplets at the beginning of the breakup calculations. The
tion to the production of too many small droplets after two unknowns, therefore, are the bigger child's droplet
separation of the parent parcel into smaller child parcels,
and thus the production of a rapid reduction in the SMR

4
radius rb and the number of drops in the child droplet ns.
Thus, Table 5. Computational Parameters for the Spray
Bomb of Alloca et al. (1992)
(Eq. 17) Number of Grid Cells 44 radial
x 149 vertical (2-D)
which is rewritten as follows: Domian Size 4.4 cm radial
x 15 cm vertical
(Eq. 18) Nozzle discharge coefficient 0.90
KH Breakup Constant (B1) 10
and the solution of this equation yields rb. The value of ns RT Breakup Constant (C3) 5.33
is easily found by solving equation (14) for ns: Temperature of All Walls in Chamber 300 K
Initial Air Temperature 300 K
Total Number of Injected Parcels 3000
Fuel Density 0.760 g/cm3
(Eq. 19) Hole Radius 100 microns
This ensures the accuracy of the solution for conserva- This implementation was chosen over an adjustment in
tion of liquid mass. the breakup time scale because the wavelength is much
longer than that of the KH model, and provides easier
IMPLEMENTATION OF RAYLEIGH-TAYLOR adjustment to spray drop sizes. By increasing the value
BREAKUP Accelerative Rayleigh-Taylor instabilities of C3, it is possible to reduce the likelihood of breakup
and aerodynamic Kelvin-Helmholtz type instabilities are and the size of the droplets after breakup are also
simultaneous phenomena in the droplet breakup process increased. Like modifying the B1 time constant of the
(Hwang et al., 1996). The models are thus allowed to Kelvin-Helmholtz model, modifying this constant allows
grow unstable waves simultaneously, and the model that an inclusion of the effects due to the initial disturbance
predicts the fastest onset of instability produces a level at the nozzle exit.
breakup event. A flowchart of the proposed breakup pro-
cess is shown in Figure 1. RESULTS
The Rayleigh-Taylor waves are only allowed to form on
the droplet when the diameter of the droplet is larger than The model is first validated using spray data from Alloca
the wavelength of the fastest growing disturbance, as et al. (1992). Table 5 shows the computational param-
predicted by equation (7). When the disturbances have eters selected. Comparisons between the predicted
grown for enough time to exceed the elapsed breakup spray behavior and the measured spray have been used
time, the droplet is split into smaller droplets which are to improve the accuracy of the model. The calculations
proportional to the wavelength of the disturbances. are all made using the standard - turbulence model.
Results using the RNG- turbulence model will be
This formulation of the Rayleigh-Taylor model reduces shown with one Caterpillar engine case.
the overall rate of division of the spray parcels, and the
resulting dispersion of droplets is also reduced. This To model the Alloca et al. (1992) case, a two-dimensional
necessitates using a prescribed cone angle to ensure cylindrical grid was used, with cells which were 1mm x
that the parcels are still dispersed within the spray cone. 1mm in a domain that was 4.4 cm in radius, and 15 cm in
This is not necessary with the wave model, since the axial length. An air temperature of 300 K and density of
breakup event is used to disperse the parcels. It also 0.0197 9/cm3 were used for the non-evaporating spray
affects the collision model, since the dispersion of parcels bomb data. Liquid penetration was calculated using the
is somewhat reduced which produces a more concen- point where 90% of the liquid mass had penetrated, simi-
trated grouping of parcels in a cell, and thereby increases lar to the experimental method.
the probability of droplet coalescence. The adjustable constants in the new RT-KH breakup
An adjustable constant, C3, was introduced to allow mod- model provided a way to control the size of the droplets
ification of the effective wavelength of the Rayleigh-Taylor emitted from the nozzle. The effect of varying the model
waves. This has the effect of changing the size of the constants on the spray penetration is shown in Figure 2.
waves and, consequently, the time between breakup Changing the KH and RT spray constants from their stan-
events. dard values of B1=10 and C3 = 5.3 to the current values
indicated has only a slight effect on the spray penetration,
while they have a large effect on droplet size, as shown in
Figure 3. The value of B1, the KH breakup delay coeffi-
cient, affects the stripping of smaller droplets from the
larger ones, so increasing its value reduces the rate of
breakup throughout the length of the spray penetration.
The value of C3 affects the larger droplets as they decel-
erate near the nozzle. A small value of C3, as seen in Fig-
ure 3, will produce a much smaller droplet size near the

5
injector. At longer distances from the nozzle, the RT comparatively small at large distances from the nozzle,
mechanism loses influence. The KH mechanism there- which in this case means distances greater than 35 mm,
fore becomes dominant, and the difference in size is even when the value of C3 is more than doubled.

Figure 1. Overall view of the Rayleigh-Taylor breakup process, as implemented in KIVA

The previous KH and new RT-KH spray models are now To show the effect of changing the breakup constants in
compared. The overall change in the predicted pene- the new RT-KH breakup model, Figure 6 illustrates the
tration is small, as seen in Figure 4, but the local droplet predictions of the new spray model with low and high val-
sizes are very different in the two models. Figure 5 shows ues of B1, the KH breakup delay coefficient. At a value of
typical results The new model consistently gives higher 10, the breakup constant only reduces the penetration of
dropsize predictions near the nozzle, and an increase in the vapor slightly compared to the predictions made with
droplet size throughout the spray. It should be noted that the constant set to 20. However, the liquid phase shows a
the experimental error, which is as high as 10 m when significant sensitivity to the breakup constant, which
the spray is dense near the nozzle, as presented in shows that B1 is effective in adjusting the liquid penetra-
Alloca et al. (1992), so it is difficult to choose among the tion length, while allowing the vapor penetration to
models. remain nearly constant.
The evaporating spray experiments of Kamimoto et al.
(1987) were used to further extend the applicability of the
spray models with the added challenge of testing the
evaporation sub-model. Table 6 shows the computational
setup for comparison with the results of Kamimoto et al.
(1987).

6
Figure 2. Effect of the spray model constants on the
penetration of the droplets Figure 4. Comparison of the previous and new models'
effect on spray penetration.

Figure 5. Comparison of the previous and new spray


breakup models' effect on droplet size at 0.3
Figure 3. Effect of the spray model constants on the
ms after the beginning of injection.
local SMR at 0.8 ms after the start of injection.
The value of the constants for the 'New Model'
The spray and vapor penetrations are more sensitive to
are B1= 10 and C3 = 5.33
the value of the RT breakup delay coefficient, C3, than to
the KH coefficient. Figure 7 shows that an increase in
Table 6. Computational Parameters for the Sprays of value of C3 from 2.5 to 5.33 (the value presented in Su et
Kamimoto et al. (1987) al (1996)), significantly increases the liquid penetration.
There is a smaller but notable effect on the vapor
Number of Grid Cells 25 radial penetration as well. The reduced breakup rate allows the
x 99 vertical (2-D)
droplets to remain larger and penetrate farther, affecting
Domain Size 3 cm radial
x 10 cm vertical both the liquid and vapor penetration. Because of the
Nozzle discharge coefficient 0.65 large sensitivity to C3, it is not possible just to choose a
KH Breakup Constant (B1) 10 value that fixes the penetration length of the vapor phase,
RT Breakup Constant (C 3) 2.5 and then adjust B1 to achieve the proper liquid phase
Temperature of All Walls in Chamber 433 K penetration.
Initial Air Temperature 900 K
Total Number of Injected Parcels 8000
Fuel Density 0.7865 g/cm3
Hole Radius 80 microns

7
spray penetration. Table 7 shows the conditions used in
the simulations

Figure 6. Effect of KH breakup constant (B1) on liquid


and vapor penetration, using the RT-KH
model. Symbols with lines represent vapor
penetration, and symbols without lines Figure 8. Liquid and vapor penetration for the RT-KH
represent liquid penetration. and KH models. Symbols with lines represent
vapor penetration, and symbols without lines
represent liquid penetration.

Figure 7. Effect of RT breakup constant (C3) on liquid


and vapor penetration, using the RT-KH
model. Symbols with lines represent vapor
penetration, and symbols without lines Figure 9. Predicted vs. measured pressure for the
represent liquid penetration. Cummins/Sandia engine of Espey et al.
(1994).
Instead, a compromise value of C3 must be chosen,
which minimizes the liquid penetration without decreas- The prediction of in-cylinder pressure is quite good, as
ing the vapor penetration greatly. shown in Figure 9. This level of pressure agreement is
comparable to that predicted by Kong, Han, and Reitz
The final comparison shows the penetration of liquid and (1995).
vapor using the new RT-KH model vs. those of the KH
model. The RT-KH model uses B1 = 10 and C3 = 2.5, and The penetration of the liquid, shown in Figure 10 was cal-
the KH model also uses B1 = 10. Figure 8 shows the liq- culated from the distance of the droplet farthest from the
uid and vapor penetration. As can be seen, the new injector in both the experiments and the predictions. Sev-
model predicts the liquid and vapor penetration much eral experimental points are given at each crank angle,
better. This combination of constants seems to work well showing the range of values observed. The present
for spray bombs under diesel-like conditions, as well as in agreement is better than that shown in Kong, Han, and
engine cases. The values may require adjustment for dif- Reitz (1995). The liquid penetration shown here is less
ferent injectors, but should not need adjustment for simi- than in the previous study, and gives a better prediction of
lar injection systems operating under similar conditions. the spray behavior. In Figure 11, the spray plots are
shown to give a physical picture of the predicted spray
The engine experiments provide less detailed spray data. behavior. Notice that the liquid penetration reaches its
The experiments of Espey et al. (1994) used a Cummins maximum length by about -4 deg. ATDC, in agreement
heavy-duty diesel engine running at low-load conditions. with the measurements of Espey and Dec. (1996).
In addition to pressure and heat release data, the trans-
parent bowl used in this engine allows observation of the

8
The Caterpillar engine is the final application for the new
Table 7. Computational parameters for the Cummins spray model. The new model correctly predicts trends for
optical-access engine. NOx and soot emissions, as well as giving good agree-
Number of Grid Cells 21 radial ment with measured cylinder pressure and heat release
x 23 azimuthal data. However, these advantages come at the cost of
x 18 vertical higher computational times. A typical calculation, with the
Domain Size 6.985 cm rad. KH model requires 8 hours on a Cray J916. The current
x 15.24 cm stroke
x 45 deg. sector
RTKH model requires 30 hours, which is mainly due to
Connecting Rod 30.5 cm the finer grid near the nozzle which is required. In return
Nozzle discharge coefficient 0.80 for that investment in time, a more accurate calculation of
Engine Speed 1200 RPM the liquid and vapor penetration results, which allows
KH Breakup Constant (B1) 10 improved calculation of emissions trends.
RT Breakup Constant (C 3) 1.0
Wall/Head/Piston Temperature 433 K
The baseline case for the Caterpillar 3406 engine is a
Initial Air Temperature 433 K single-injection at -11 degrees ATDC. The engine details
Intake Valve Close -150 deg. ATDC are given in Table 8 and Figure 12. To give the best com-
Total Number of Injected Parcels 1000 parison between the models, the spray constant, B1, was
Fuel Density 0.7865 g/cm3 set to a value of 30.
Fuel Mass per Injection 0.0550 g
Initial Fuel Temperature 340 K
Compared to the new RT-KH model, the previous KH
Hole Radius 97.0 microns model, optimized for combustion, gives a slow initial pres-
sure rise, and an earlier peak, as seen in Figure 13. The
Table 8. Computational Parameters for the Caterpillar results for both models are very close to the measure-
engine ments, but produce very different results for the emis-
sions, as will be seen. The heat release, seen in Figure
Number of Grid Cells 29 radial
x 20 azimuthal 14, from the previous model is lower for the premixed
x 35 vertical burn, and occurs slightly too early. At the same time, the
Domain Size 6.858 cm rad. premixed heat release for the new model is seen to be
x 16.51 cm stroke slightly too high. The diffusion burn predicted by the pre-
x 60 deg. sector
vious model is likewise too early and slightly too slow.
Connecting Rod 26.3 cm
Nozzle discharge coefficient 0.8
Engine Speed 1600 RPM
KH Breakup Constant (B1) 10
RT Breakup Constant (C 3) 2.5
Wall Temperature 433 K
Head Temperature 523 K
Piston Temperature 553 K
Initial Air Temperature 342 K
Intake Valve Close -147 deg. ATDC
Total Number of Injected Parcels 1500
Fuel Density 0.720 g/cm3
Initial Fuel Temperature 341 K
Hole Radius 129.5 microns

Figure 10. Predicted vs. measured liquid penetration for


the Cummins/Sandia engine of Espey et al.
(1994)

9
Figure 11. Spray plots for the Cummins/Sandia engine from -8 deg ATDC to TDC

The fuel injected, evaporated, and burned, presented in


Figure 15, show that the previous model evaporates the
fuel too slowly, while allowing the fuel vapor distribution in
the engine to be such that it ignites quickly, burning all of
the available fuel between -5 and -3 deg. ATDC. This
accounts for the weak premixed burn. In the short time
span between just before TDC and approximately 1 deg.
ATDC, the evaporation rate climbs dramatically, helped Figure 12. Computational grid in the plane of the spray in
by the early combustion. Between 7 and 15 deg. ATDC, the Caterpillar engine. The azimuthal sector is
the rate of combustion slows as the fuel dispersion in the 60 degrees with cells placed at equal
engine is dominated by large droplets. The end result is increments of 3 degrees.
that the pressure shows reasonable agreement with
experiments, but the details of the breakup, evaporation,
and combustion processes show the improvements
made by the new model.

10
The emissions predictions show that soot is formed later
and remains higher for most of the combustion process
with the previous model, as seen in Figure 16. The rea-
son for this is the tendency of the previous model to
maintain a long column of liquid fuel which forms a rich
region around it, causing more soot to be observed
between 20 and 60 degrees. Late in the cycle, the soot
oxidizes more quickly, due to the higher penetration of the
liquid phase. The NOx prediction, on the other hand, is
higher at all times with the previous KH model, as shown
in Figure 17.
To allow assessment of the emissions trends, the emis-
sions predictions are multiplied by constants to repro-
Figure 13. Cylinder pressure in the Caterpillar engine duce agreement with the measured value for the baseline
with injection at -8.5 deg. ATDC, shown for the case. For the NOx predictions, the results are multiplied
new spray model vs. the previous model with by 2.05, and for soot, the results are multiplied by 3.89 to
B1 = 30 match the experiments. These values are kept constant
throughout the remainder of this study. In the compari-
sons with early injection and split injections which will be
presented later, the current model produces good agree-
ment with the experiments using these constants. In the
comparison given here, the current model produces bet-
ter agreement in the details of the combustion process as
well as the pollutant production process.
The turbulence model has a strong effect on the com-
bustion and, in particular, the emissions prediction. The
RNG model predicts a strong increase in the penetration
of the fuel into the combustion chamber. This results in a
very focused spray and a vapor cone that penetrates fur-
ther into the engine, but it creates a low temperature
zone around the liquid fuel which does not allow enough
vapor to be produced early in the combustion. The pres-
Figure 14. Heat release in the Caterpillar engine with sure is therefore low with the same spray model con-
injection at -8.5 deg. ATDC, shown for the new stants as used for the - model, as seen in Figure 18.
spray model vs. the previous model with B1 = The RNG model, however, does seem to give a qualita-
30 tively better shape to the heat release curve, and as
shown in Figure 19, which has a lower initial peak, a
stronger drop-off of the initial heat release, consistent
with the measured curve, and only a slightly higher diffu-
sion burn peak, which be gins too early, but otherwise
matches the behavior of the - model well after 10 deg.
ATDC. The differences in the combustion behavior are
also, in part, due to the change in model constants used
in conjunction with the RNG-kc model. The constants
used in both turbulence models are as presented in
Kong, Han, and Reitz (1995). This means C=0.0845,
C1=1.42, C2=1-68, = = 1.39, 0=4.38, and
=0.012. The constants used with the standard -
model are C=0.09, C1=1.44, C2=1.92, =1-0, and
= 0.769.
Figure 15. Cumulative percentage of fuel injected,
evaporated, and burned in the Caterpillar
engine with injection at -8.5 deg. ATDC, shown
for the new spray model vs. the previous
model with B1 = 30

11
Figure 19. Heat release comparison between the RNG
and - models in the Caterpillar engine with
Figure 16. Calculated soot emissions using the new
injection at -8.5 deg. ATDC.
RTKH and previous KH spray model with B1 =
30 in the Caterpillar engine with injection at
-8.5 deg. ATDC.

Figure 20. Cumulative percentage of fuel injected,


evaporated, and burned in the Caterpillar
engine with injection at -8.5 deg. ATDC using
Figure 17. Calculated NOx emissions using the new
the RNG and - turbulence models
spray model and the previous spray model
with B1 = 30 in the Caterpillar engine with
injection at -8.5 deg. ATDC

Figure 18. Cylinder pressure comparison between the


RNG and - models in the Caterpillar engine
with injection at -8.5 deg. ATDC.

12
Figure 22. Cylinder pressure in the Caterpillar engine
with injection at -2.5 deg. ATDC.

Figure 21. Calculated soot and NOx emissions


(uncorrected) using the new spray model in
the Caterpillar engine with injection at -8.5
deg. ATDC with the RNG and - turbulence
models.
Figure 23. Heat release in the Caterpillar engine with
The fuel injected, evaporated, and burned is shown in injection at -2.5 deg. ATDC.
Figure 20. The evaporated fuel in the RNG model is lower
than that of the - model, and consequently the rate of The - turbulence model was used for the present
combustion is also lower. The effect on the soot and NOx study, with the multiplication factors described previously
predictions is shown in Figure 21 (neither plot uses the for the soot and NOx emissions applied to the emissions
correction factor explained previously, to illustrate the output. In future studies, it is expected that the RNG
effect of the turbulence model.) The soot prediction is model will produce slightly better results, based on the
insensitive to the turbulence model at the end of the pre- shape of the heat release curve, and the study of Han
diction. The NOx prediction is nearly three times higher and Reitz (1995). However, the turbulence model only
with the RNG model, again indicating that the diffusion slightly changes the overall spray and vapor behavior
burn is more localized and at a higher temperature than with the new spray model, and while the magnitudes of
that of the - model. This agrees with the findings of the emissions show the effect of the turbulence model,
Han and Reitz (1995). the emissions trends are relatively insensitive to the tur-
bulence model used.
The case of late injection is a challenging case for mod-
eling. Figure 22 shows that the new model predicts the
shape of the pressure curve correctly. The premixed burn
is somewhat too strong, as seen in the heat release
curve of Figure 23, but the diffusion burn is at the correct
magnitude. While the overall prediction made by the new
spray model is good, it appears that the behavior of the
model may need to be refined further at the very begin-
ning of the injection.

13
The cumulative injected, evaporated, and burned fuel
masses show that the fuel from the initial injection is
almost completely burned before the start of the second
injection. In this case, the limit is good because the diffu-
sion burn is too strong at the beginning. The second dif-
fusion burn seems to progress at the correct rate,
however, the combustion begins without delay as soon as
more fuel vapor is available. The good overall agreement
with the combustion process leads to good agreement in
the emissions calculations, as seen in Figure 25 (b).
The importance of splitting the injection is seen in the
soot curve. The interruption between injections allows the
soot to oxidize in that period, and this reduces the peak
Figure 24. Soot and NOx emissions in the Caterpillar value of soot observed within the cylinder (Han et al.,
engine with injection at -2.5 deg. ATDC. 1996). Because the combustion is interrupted, the tem-
peratures are also kept lower, and the NOx level is only
The ignition event is correctly timed at 3 deg. ATDC, as slightly higher than that of the single injection case,
seen in Figure 23. This ignition event creates a strong despite starting 2.5 deg. earlier in the cycle, indicating
increase in the vaporization rate, and leads to the strong that a new tradeoff level has been reached.
premixed burn. Because more fuel is consumed early in The final case to be considered is the 75-(8)-25 injection
the cycle, not enough fuel is left to sustain the stronger, strategy. In this case, there is good pressure agreement,
hotter diffusion burn observed in the experiments, and as seen in Figure 25 (c), as well as good agreement in
this leads to an under prediction of the NOx, shown in the heat release rate. With the exception of the premixed
Figure 24. The soot prediction is lower than the experi- burn, the agreement in the heat release curve is excel-
mental results. lent.
The final topic of this study is the prediction of split injec- Once again, the cumulative injected, evaporated, and
tion cases. The three cases chosen for study have the burned fuel mass indicates that the dwell between injec-
same dwell of 8 degrees between the end of the first tions is long enough to interrupt the combustion process.
injection and the beginning of the second injection. An In this case, the effect on the emissions is easily visible in
injection with half of the total spray mass in each of the Figure 25 (c). The rate of increase in NOx is slowed
two injections with an 8 deg. dwell between injections significantly, and this happens just at the time when the
would be listed as 50-(8)-50. The injection strategies cho- rate of NOx production is the fastest, and the rate of pro-
sen for this study had 25, 50, and 75% of the fuel injected duction does not return to its original strength. This inter-
in the first spray pulse (Nehmer, 1993). ruption in the NOx formation process, combined with a
The first case to be considered is the 25-(8)-75 injection lengthened period of soot oxidation shows why this strat-
strategy. The pressure curve shows reasonably good egy has had success in other studies on this engine (Tow,
agreement with measurements in Figure 25 (a) The heat Pierpont, and Reitz, 1994).
release rate indicates the rate of combustion drops too
quickly near TDC, and the diffusion burn is slower and CONCLUSIONS
much less intense than the experiments indicate.
A new spray model has been introduced and its
The combustion rate is limited by the fact that injection
implementation has been documented. This new spray
stops near TDC, The first injection ends at approximately
model accounts for the physical phenomena of accelera-
-4 deg. ATDC, which is also the point at which the pres-
tive, Rayleigh-Taylor type instabilities on the surface of a
sure begins to deviate from the experimental value.
droplet, as well as accounting for the Kelvin-Helmholtz
Nearly all of the fuel from the first injection is consumed waves which form due to aerodynamic instabilities. The
before the second injection begins, and the rate of com- simulation results indicate that the new spray model,
bustion also slows dramatically. The effect upon engine applied to a heavy-duty diesel engine, shows improved
emissions is that NOx is low, as seen in Figure 25 (a), combustion predictions and a spray that matches,
due to the slow combustion rate, but the soot prediction is qualitatively, the short liquid penetration lengths observed
reasonably accurate, due to accurate prediction of the optically in a diesel engine operating under similar condi-
length of the combustion event and the amount of oxida- tions (Espey et al., 1994). The model shows very good
tion that took place. agreement with early injection cases, and good overall
agreement in combustion behavior for all of the cases
The second split injection case, the 50-(8)-50 strategy,
presented here. Compared to previous results (e.g.
shows slightly better pressure agreement when com-
Patterson et al., 1994), the emissions predictions for the
pared to experiments. The pressure, shown in Figure 25
split injection strategies were much more accurate, and
(b) shows good agreement with the experiments.
represent the observed emissions trends quite well. The

14
previous model overestimates the amount of NOx as well
as giving a poorer trend in the soot-NOx tradeoff curve.
The model also indicates that a successful injection strat-
egy uses the dwell between injections to utilize all of the
air in the combustion chamber more effectively. This
results in lower peak temperatures combined with a
longer soot oxidation process that allows both emissions
to be reduced effectively. The ability to determine the
mechanism of emissions formation can now be used to
further investigate which injection strategies are the most
effective at reducing emissions while maintaining engine
power and driveability. In summary, the following points
were demonstrated:
The new spray model provides a more accurate cal-
culation of the local droplet size, compared to exper-
imental data.
Prediction of the liquid and vapor phases of fuel pen-
etration is improved, both in evaporating spray
bombs, and in engines.
The new model reduces the amount of liquid-wall
impingement in the Caterpillar engine, which agrees,
qualitatively, with experimental data.
Combustion and heat release are predicted more
accurately than in previous studies, particularly for
the split injections.
Trends in the emissions of soot and NOx are now
more accurately represented, allowing further study
with more complex injection strategies.

(a)

15
(b) (c)

Figure 25. Cylinder pressure (top row), heat release (second row), cumulative fuel mass injected, evaporated, and burned
(third row), and Soot-NOx emissions (fourth row) in the Caterpillar engine for three split injection schemes. Scheme (a) is
25-(8)-75 with start of injection at -13 deg. ATDC, scheme (b) is a 50-(8)-50 with start of injection at -11 deg. ATDC, and
scheme (c) is a 75-(8)-25 with start of injection at -11 deg. ATDC.

16
ACKNOWLEDGMENTS 20. Su, T. F., Patterson, M. A., Reitz, R. D., and Farrell, P. V.,
"Experimental and Numerical Studies of High Pressure
Multiple Injection Sprays", SAE Paper 960861, 1996.
The authors would like to acknowledge financial support 21. Taylor, G. I., "The Instability of Liquid Surfaces When
given by Cray Research, NASA-DOE grant NAG3-1087, Accelerated in a Direction Perpendicular to Their Planes,"
TACOM, Caterpillar, and the Army Research Office. The Scientific Papers of G. I. Taylor, ed. G. K. Batchelor,
Vol. III, University Press, Cambridge, 1963.
REFERENCES 22. Tow, T., Pierpont, A., and Reitz, R. D., "Reducing Paxticu-
lates and NOx Emissions by Using Multiple Injections in a
Heavy Duty D.I. Diesel Engine", SAE Paper 940897, 1994.
1. Alloca, L., Belardini, P., Bertoli, C., Corcione, F., and De
Angelis, F. "Experimental and Numerical Analysis of a Die-
sel Spray" SAE Paper 920576, 1992.
2. Amsden, A.A., O'Rourke, P.J. and Butler, T.D., "KIVA-II - A
Computer Program for Chemically Reactive Flows with
Sprays," Los Alamos National Labs., LA-11560-MS, 1989.
3. Bellman, R. and Pennington, R. H., "Effects of Surface Ten-
sion and Viscosity on Taylor Instability," Quarterly of
Applied Mathematics, Vol. 12, 1954.
4. Chang, S. K., "Hydrodynamics of Liquid Jet Sprays," Ph.D.
Thesis, University of WisconsinMadison, 1991.
5. Espey, C., Dec, J. E., Litzinger, T. A., and Santavicca, D. A.,
"Quantitative 2-D Fuel Vapor Concentration Imaging in a
Firing D. I. Diesel Engine Using Planar Laser-Induced Ray-
leigh Scattering," SAE Paper 940682, 1994.
6. Han, Z. Y., Uludogan, A., Hampson, G. J., and Reitz, R. D.,
"Mechanism of Soot and NOx Emission Reduction Using
Multiple-Injection in a Diesel Engine" SAE Paper 960633,
1996.
7. Han, Z. Y., and Reitz, R. D. "Turbulence Modeling of Inter-
nal Combustion Engines using RNG - Models," Com-
bust. Sci. and Tech., vol. 106, pp. 267-295,1995.
8. Hwang, S. S., Liu, Z., and Reitz, R. D., "Breakup Mecha-
nisms and Drag Coefficients of High Speed Vaporizing Liq-
uid Drops," Atomization and Sprays, vol. 6, pp. 353-376,
1996.
9. Kamimoto, T., Yokota, H., and Kobayashi, H., "Effect of
High Pressure Injection on Soot Formation Processes in a
Rapid Compression Machine to Simulate Diesel Flames,"
SAE Paper 871610, 1987.
10. Kong, S. C., Han, Z., and Reitz, R. D., "The Development
and Application of a Diesel Ignition and Combustion Model
for Multidimensional Engine Simulations," SAE Paper
950278, 1995.
11. Miyaki, M., Fujisawa, H., Masuda, A., and Yamamoto, Y.,
"Development of New Electronically Controlled Fuel Injec-
tion System ECD-U2 for Diesel Engines," SAE Paper
910252, 1991.
12. Nehmer, D. A., "Measurement of the Effect of Injection
Rate And Split Injection on Diesel Engine Soot and NOx
Emissions," MS Thesis, University of Wisconsin-Madison,
1993.
13. 0' Rourke, P. J. and Amsden, A. A., "The TAB Method for
Numerical Calculation of Spray Droplet Breakup," SAE
Paper 872089, 1987.
14. Patterson, M. A., Kong, S. C., Hampson, G. J., and Reitz,
R. D., "Modeling the Effects of Fuel Injection Characteris-
tics on Diesel Engine Soot and NOx Emissions" SAE Paper
940523, 1994.
15. Reitz, R. D. "Modeling Atornization Processes in
High-Pressure Vaporizing Sprays," Atomization and Spray
Technology, 3, 309-337, 1987.
16. Reitz, R. D. and Diwakar, R., "Structure of High-Pressure
Fuel Sprays" SAE Paper 870598, 1987.
17. Reitz, R. D., "Computer Modeling of Sprays," Spray Tech-
nology Short Course, Pittsburgh, PA, May 17, 1994.
18. Reitz, R. D. and Bracco, F. V., "Mechanisms of Breakup of
Round Liquid Jets," Encyclopedia of Fluid Mechanics, Ch.
10, Gulf Publishing Company, Houston, 1986.
19. Su, T. F. "An Experimental Study of High Injection Pressure
Diesel Sprays", Ph.D. Thesis, University of Wisconsin-Mad-
ison, 1995.

17

You might also like