You are on page 1of 6

Comparison of Characteristic Time (CTC), Representative Interactive Flamelet (RIF), and

Direct Integration with Detailed Chemistry Combustion Models against Multi-Mode


Combustion in a Heavy-Duty, DI Diesel Engine
Satbir Singh and Rolf D. Reitz
Engine Research Center, Department of Mechanical Engineering, University of Wisconsin, Madison
Abstract
Three different approaches for modeling diesel engine combustion are compared against cylinder pressure, NOx emissions, and
simultaneous optical diagnostic images from a heavy-duty, DI diesel engine. A characteristic time combustion (KIVA-CTC) model,
a representative interactive flamelet (KIVA-RIF) model, and direct integration using detailed chemistry (KIVA-CHEMKIN) were
integrated into the same version of the KIVA-3v computer code. In this way, the computer code provides a common platform for
comparing various combustion models. Five different engine operating strategies that are representative of several different
combustion regimes were explored in the experiments and model simulations.
Comparison of simulated cylinder pressure and heat-release rates with the experimental results shows that all the models predict the
cylinder pressure and heat release rate reasonably well. All the models predict NOx emissions trends very well but the absolute
magnitudes are different from the experimental measurements. A comparison of the model predicted and experimentally observed
in-cylinder details is also presented. The KIVA-CHEMKIN model better predicted the details of the flame structure and emissions
for the LTC conditions.

INTRODUCTION

This paper presents a comparison of three previously proposed


combustion models with experimental data. These models are
the characteristic time combustion (KIVA-CTC), CFD
integrated with detailed chemistry (KIVA-CHEMKIN), and
the representative interactive flamelet (KIVA-RIF) models. In
order to test the predictive capability of these models in
different types of combustion environments, a set of
experiments that belong to five different combustion strategies
was designed. All three models are first validated against the
measured cylinder pressure, heat release data and NOx
emissions and a comparison with in-cylinder images of
combustion is also presented.

One significant benefit of computational models is that they


can reduce engine research and development time and
resources through their predictive capabilities. They also
complement experimental observations by providing
information not available from the experiments [1]. Model
predictions are reliable over a limited range of operating
conditions, which depends upon the complexity of the model
and the experimental data against which it was calibrated. As
the engine community is exploring new engine technologies to
enhance fuel conversion efficiency and reduce pollutant
emissions, computer models are being challenged to perform
over a wider range of engine operating conditions. These
operating conditions range from conventional, hightemperature, heterogeneous diesel diffusion combustion to
more homogeneous low-temperature combustion (LTC). This
requires high-fidelity models that can cover an expanded
range of combustion environments. In addition to maintaining
their good predictive capability, short calculation times are
also desirable.

EXPERIMENTAL SETUP
ENGINE AND OPERATING CONDITIONS
A single-cylinder, direct-injection (DI), 4-stroke diesel engine
based on a Cummins N-series production engine was used in
this investigation. The engine has a bore of 139.7 mm and
stroke of 152.4 mm. A schematic diagram of the engine is
shown in Fig. 1, and the complete description of the engine is
available in Refs. [5,6]. The engine is equipped with a nonproduction, high-pressure, electronically-controlled, commonrail fuel injector. Specifications for the fuel injector are also
available in Refs. [5,6]. The fuel used is EPA certified 2007
diesel fuel.

Multi-dimensional computational fluid dynamics (CFD)


models have contributed to various aspects of engine
development in recent years, with practical computational
cost [2,3]. For example, a multi-dimensional computer code
KIVA-3v coupled with a genetic algorithm was recently used
by Wickman et al. [4] to optimize the combustion chamber
bowl geometry for reduced emissions and improved fuel
conversion efficiency. The predictive capability of multidimensional models is limited by the accuracy of their submodels, the most important of which are the turbulence, spray
and combustion models.

Five different operating conditions were selected to be


representatives of five modes of diesel combustion, as
described in Table 1. Note that the intake-temperatures and
pressures in Table 1 are significantly higher than typical for
production engines because of the low compression ratio of
the optical engine.
1

Table 1 Test Matrix

Speed (RPM)
IMEP (bar)
Injection Pressure (bar)
Intake Temp (C)
BDC Temp (C)
Intake Pressure (kPa)
TDC Motored Temp. (K)
TDC Motored Density (kg/m3)
Peak Adiabatic Flame Temp. [K]
SOI (ATDC)
Injection Quantity (mg)
DOI (CAD)
O2 Conc. (Vol %)

High-T
Short-ID
1200
4.4
1200
111
106
233
905
24
2760
-7
61
10
21

High-T,
Long-ID
1200
4.5
1200
47
62
192
800
22.3
2700
-5
61
10
21

Low-T,
Early-Inj.
1200
3.9
1600
90
92
214
870
22.9
2256
-22
56
7
12.6

Low-T,
Late-Inj.
1200
4.1
1600
70
78
202
840
22.5
2164
0
56
7
12.6

Low-T,
Double-Inj.
1200
4.5
1600
90
92
214
870
22.9
2132
-22, + 15
31, 33
4, 4
12.6

conditions, and are therefore characterized as LTC conditions.

The first two conditions in Table 1 are characterized as hightemperature because the charge gas is air (21% oxygen by
volume), and the resulting stoichiometric flame temperatures
are relatively high.

OPTICAL DIAGNOSTIC TECHNIQUES


A number of optical diagnostics were applied to investigate
various in-cylinder processes of spray, combustion and
emissions for operating conditions described in Table 1. A
brief summary of these diagnostics is given in Table 2 and a
detailed description can be found in references given next to
each diagnostic.

For the three remaining conditions, the oxygen volume


fraction was reduced to 12.7% via nitrogen dilution. Because
of the dilution effect of the added nitrogen, the adiabatic,
stoichiometric flame temperatures for the three conditions
with 12.7% oxygen are much lower than for the two undiluted

COMPUTATIONAL MODELS
The simulations were performed using the KIVA-3v release 2
code implemented with three different combustion models:
KIVA-CTC [7], KIVA-CHEMKIN [8], and KIVA-RIF [9].
The hybrid Kelvin Helmholtz Rayleigh Taylor (KH-RT)
model [10] predicted the spray breakup and the RNG- k-
model developed by Han and Reitz [11] predicted the
turbulence.
The mesh was composed of about 80,000 computational cells
at bottom dead center (BDC) with 1.2 x 1.2 x 1.2 mm cell size
Table 2 Optical Diagnostics Summary
Diagnostic
Laser-elastic (Mie)
scattering [12]
OH PLIF [13]
Broadband (BB) PLIF [14]

Soot PLII [15,16,17]


Chemiluminescence
[16,18]
Soot luminosity [16,19]

Figure 1. Experimental setup

Measurement
Planar liquid-fuel spray
imaging (LL)
Planar imaging of OH
radicals
Planar imaging of
unburned fuel and
combustion intermediates
Planar soot distribution
Line-of-sight ignition
locations
Line-of sight hot soot
distribution

in the bowl near TDC for adequate resolution.

the KIVA-CHEMKIN model gives a slightly better prediction


of NOx emissions.

RESULTS AND DISCUSSION

MODEL COMPARISON WITH IN-CYLINDER IMAGES


HEAT RELEASE RATE, CYLINDER PRESSURE AND
NITROGEN OXIDE (NOX) EMISSIONS

The experimental investigation provided in-cylinder details of


spray, auto-ignition sites, OH radicals, and soot distributions
in the jet for all five operating conditions. However, in this
paper, only a very brief comparison of the model predictions
with the experimental observations is made. Additional
detailed comparisons can be found in Refs. [20,21].

Figures 2(a) to 2(e) show comparisons of the measured and


the model predicted cylinder pressure and AHRR for all five
operating conditions. For the KIVA-RIF model, the results
are presented for a single flamelet (RIF_1) in the
computational domain. For the high-T, short-ID condition
(Fig. 2a, all the models predicted the ignition delay reasonably
well but the peak cylinder pressure is over predicted. For the
high-T, long-ID condition (Fig. 2b), KIVA-CHEMKIN and
KIVA-RIF predicted a shorter ID, and KIVA-CTC predicted a
longer ID than the experiment. The KIVA-RIF model
predicted very high peak AHRR as compared to the KIVACTC and the KIVA-CHEMKIN models.

1. High-T, Long-ID Condition


Figure 3 shows comparisons of experimentally observed OH
PLIF (green) and the line-of-sight soot luminosity (red) with
the model predicted OH (green) and soot (red) in a plane along
the fuel jet axis for the high-T, long-ID condition (Table 1).
Since the KIVA-CTC model does not predict OH, only the
soot distribution is shown. Only one image at the time of peak
AHRR is presented from the model predictions and the
experimental observations (Figure 2b).

For the three low-T conditions (Figs. 2c,d, and e), the cool
flame (CF) ignition delay is under-predicted by the KIVACHEMKIN and the KIVA-RIF models. The KIVA-CTC
model that uses the simplified SHELL model for predicting
auto-ignition, does not predict a distinct CF heat release. For
the low-T, early-Inj. (Fig. 2c), and the first combustion event
of low-T, double-Inj condition (Fig. 2e), all the models
predicted the second stage combustion (SSC) ignition delay
accurately, but for the low-T, late-Inj. condition (Fig. 2d), the
KIVA-RIF model overpredicted and the KIVA-CHEMKIN
model slightly underpredicted the ignition delay. The peak
AHRR prediction is different for each model. Generally, the
KIVA-RIF and the KIVA-CTC models predicted faster rate of
combustion and higher peak AHRR. For the KIVA-CTC
model, the standard ignition and combustion model constants
[7] were slightly tuned to get a better prediction of the ID and
rate of combustion, but they were kept the same for all of the
operating conditions.

The experimental image in the left column shows a high


concentration OH ring at the jet periphery surrounding a high
soot concentration region. This distribution of OH and soot is
very similar to the conceptual model proposed by Dec [5].
The KIVA-CTC model that uses fuel as soot pre-cursor,
predicts soot throughout the jet. The soot model is activated
when the temperature of the computational cell approaches
1100 K, and the local soot concentration is proportional to the
amount of fuel in the cell. Comparison with the experimental
images shows that the KIVA-CTC model predicts soot over a
larger region of the jet in disagreement with the experiment.
Also, the model predicts soot in the upstream regions of the jet
long after the end of combustion. This is also clearly in
disagreement with the experimental observations.
Both the KIVA-CHEMKIN and the KIVA-RIF models predict
a diffusion flame surrounding the soot-producing region. The
predicted flame structures are somewhat different for the two
models. The KIVA-RIF model predicted a broader OH
distribution and the flame extends all the way back to the
injector while the KIVA-CHEMKIN model predicted a
thinner diffusion flame that terminates some distance away
from the injector in agreement with the experiments. The
experimental image shows some weak fluorescence in the near
nozzle area but it is believed that this fluorescence is from
unburned fuel, and OH is not present near the injector. For the
KIVA-RIF model, a large soot producing region overlaps with
the OH regions as indicated by the light yellow colored
regions of the image. On the other hand for the KIVACHEMKIN model, the soot and OH regions are spatially
displaced, as also seen in the experiments.

Figure 2(f) shows the measured and predicted nitric oxide


(NOx) emissions for all the five operating conditions. The
30 ppm and 40 ppm horizontal lines on Figure 2(f)
identify a break in the vertical scale. The scale for the low
NOx levels is set from 0-30 ppm and for the high levels, it is
set from 40-1000 ppm. The two high-temperature operating
conditions have high NOx emissions, while the three lowtemperature conditions have very low NOx emissions. The
KIVA-CHEMKIN model over-predicts NOx emissions for the
high-T, short-ID condition due to the over-prediction of
cylinder pressure. Although the KIVA-RIF and the KIVACTC models also over-predicted the cylinder pressure, due to
their low in-cylinder peak temperature prediction [20], the
NOx emissions are lower than the KIVA-CHEMKIN
prediction and compare well with the experimental
measurements. Similarly, for the high-T, long-ID condition,
KIVA-CHEMKIN predicted NOx emissions are higher than
the KIVA-RIF and the KIVA-CTC predictions, but they match
better with the experiment. For the three low-T conditions,
3

High-T, Short-ID

TEST
CTC
CHEMKIN
RIF_1

7
1200

6
5

400
4
300
3
200

Injection Profile

100

-100
-20

-15

-10

-5

10

15

20

7
6
5

900
4
600

3
Injection Profile
2

300

1
0

-1
25

-15

-10

-5

1100
TEST
CTC
CHEMKIN
RIF_1

800

700

600

500

500

4
3

Injection Profile

300

200
100
0

AHRR (J/deg)

AHRR (J/deg)

900

600

Low-T, Late-Inj.

TEST
CTC
CHEMKIN
RIF_1

8
7

400

300

Injection Profile

100

2
1

-25

-20

-15

-10

-5

10

-100
-10

-1
15

-5

10

25

30

1000

Low-T, Double-Inj.

TEST
CTC
CHEMKIN
RIF_1

600
500

600

4
3
2

100

NOx (ppm)

300

400

Pressure (MPa)

Injection Profile

TEST
CTC
CHEMKIN
RIF_1

800
7

400

-100
-30

20

2(d)

2(c)

200

15

CAD ATDC

CAD ATDC

AHRR (J/deg)

200

-100
-30

700

25

8
Low-T, Early-Inj.

700

400

20

Pressure (MPa)

800

15

2(b)

2(a)

900

10

CAD ATDC

CAD ATDC

1000

Pressure (MPa)

AHRR (J/deg)

500

High-T, Long-ID

TEST
CTC
CHEMKIN
RIF_1

1500

AHRR (J/deg)

600

Pressure (MPa)

700

Pressure (MPa)

800

200
40 ppm
30 ppm

25
20
15
10
5
0

-20

-10

10

20

CAD ATDC

2(e)

30

40

50

High-T
Short ID

High-T
Long ID

Low-T
Early Inj.

Low-T
Late Inj.

Low-T
Double Inj.

2(f)

Figure 2. Cylinder pressure, AHRR, and NOx emissions for all the five operating conditions

Figure 3. Comparison of the experimentally observed OH PLIF (green) and soot luminosity (red) with the
model predicted OH (green) and soot (red). The white dot donates the injector location
4

Figure 4. Comparison of the experimentally observed liquid fuel (blue) and ignition chemiluminescence
(green) with the model predicted liquid fuel (blue) and gas temperature (green)

Figure 5. Comparison of the experimentally observed OH PLIF (green) and soot LII (red) with the model
predicted OH (green) and soot (red)

CONCLUSIONS
2. Low-T, Early-Injection Condition

Three different approaches of modeling ignition and


combustion in diesel engines were compared against multimode combustion experiments on a heavy-duty DI diesel
engine. The models were implemented into the same version
of KIVA-3v to provide a common platform for comparing
previously proposed ignition and combustion models. Five
different engine operating strategies that are representative of
several different combustion regimes were explored in the
experiments and model simulations.
The following
conclusions can be drawn based on the observations:

Figure 4 shows simultaneous images of the experimentally


observed liquid-fuel (blue) and naturally occurring luminous
emission from chemiluminescence (green) and the model
predicted liquid-fuel (blue color and red particles) and local
gas temperature (green) in a plane along the fuel jet axis. The
images are presented at the peak of cool flame (CF) heat
release (Fig. 2c) and therefore, the temperature scale for the
model images is set from 850 1100 K only.
At the peak of the CF heat release, the experimental image
shows strong chemiluminescence from mid-stream to
downstream regions of the jet. The KIVA-CHEMKIN and the
KIVA-RIF models also predict CF ignition in similar regions
of the jet; however, the actual region of high temperature is
different for the two models. The KIVA-RIF model also
predicts ignition near the injector tip, in disagreement with the
experimental image.
The comparison of the model predicted OH and soot
distributions with the experimentally observed OH via OH
PLIF and soot via soot LII is also given in Figure 5. The
results show that both the KIVA-CHEMKIN and the KIVARIF models predict the OH distributions very well. Again, in
the experimental image, the fluorescence in the upstream
regions is believed to be from unburned or partially burned
fuel. Similar to the experiments, the KIVA-CHEMKIN and
KIVA-RIF models predict soot near the bowl edge but again,
for the KIVA RIF model, some of the soot is produced in
regions that overlap with the OH regions (yellow color).

1.

All three combustion models give reasonable


predictions of the cylinder pressure and heat release
rate trends for all five combustion strategies modeled.

2.

The NOx emission predictions by KIVA-CHEMKIN


follow the cylinder pressure prediction, i.e., overprediction of cylinder pressure results in overprediction of NOx emissions (e.g., the high-T,
short-ID condition). On the other hand, KIVA-RIF
tends to under-predict NOx emissions. This is
thought to be due to the use of a global scalar
dissipation rate, which introduces excessive
averaging into the physical domain. In spite of
simplified representation of fuel chemistry, the
KIVA-CTC model predicts trends in cylinder
pressure and NOx emissions reasonably well.

3.

The KIVA-CHEMKIN and the KIVA-RIF predicted


OH and soot distributions are similar to the
experimental observations, however, the flame
structure predicted by the KIVA-RIF model is more
smeared out. Also, the KIVA-RIF model predicts

11. Han, Z. Y. and Reitz, R. D., "Turbulence Modeling of


Internal Combustion Engines using RNG k-e Models,"
Comb. Sci. Tech., 106, 267, 1995.

ignition close to the nozzle, which was not observed


in the experiments.

REFERENCES
1.

2.

3.

4.

5.

14. Musculus, M.P.B. Multiple Simultaneous Optical


Diagnostic Imaging of Early-Injection Low-Temperature
Combustion in a Heavy-Duty Diesel Engine, SAE Paper
06P-72, 2006.

Reitz, R. D. and DiWakar, R., "Effect of Drop Breakup


on Fuel Sprays," SAE Paper 860469, SAE Transactions,
95, No. 3, pp. 218-227, 1986.

15. Dec, J.E. Soot Distribution in a D.I. Diesel Engine using


2-D Imaging of Laser-Induced Incandescence, Elastic
Scattering, and Flame Luminosity, SAE Paper 920115,
SAE Transactions, 101(4), 101-112, 1992.

Wickman, D. D., Senecal, P. K., and Reitz, R. D., "Diesel


Engine Combustion Chamber Geometry Optimization
using Genetic Algorithms and Multi-Dimensional Spray
and Combustion Modeling," SAE Paper 2001-01-0574,
2001.

16. Dec, J. E. and Espey C. Ignition and Early Soot


Formation in a D.I. Diesel Engine using Multiple 2-D
Imaging Diagnostics, SAE Paper 950456, SAE
Transactions, 104(3), 1995.

Dec, J. E., "A Conceptual Model of D.I. Diesel


Combustion Based on Laser-Sheet Imaging," SAE Paper
970873, SAE Transactions, 106, No. 3, pp. 1319-1348,
1997.
Musculus, M. P. B., "On the Correlation between NOx
Emissions and the Diesel Premixed Burn," SAE Paper
2004-01-1401, SAE Transactions, 113, No. 4, pp. 13191348, 2004.

7.

Kong, S-C., Han, Z. Y., and Reitz, R. D., "The


Development and Application of a Diesel Ignition and
Combustion Model for Multidimensional Engine
Simulations," SAE Paper 950278, SAE Transactions, 104,
No. 3, pp. 502-518, 1995.

9.

13. Dec, J.E. and Coy, E.B. OH Radical Imaging in a DI


Diesel Engine and the Structure of Early Diffusion Flame.
SAE Paper 960831, SAE Transactions, 105(3), 11271148, 1996.

O'Rourke, P. J. and Amsden, A. A., "Three Dimensional


Numerical Simulations of the UPS-292 Stratified Charge
Engine," SAE Paper 870597, SAE Transactions, 96, No.
4, pp. 710-727, 1987.

6.

8.

12. Espey, C. and Dec, J.E. The effect of TDC Temperature


and Density on the Liquid-Phase Penetration in a D.I.
Diesel Engine. SAE Paper 952456, SAE Transactions,
104(4), 1400-1414, 1995.

Miles, P., Choi, D., Megerle, M., RempelEwert, B., Reitz,


R. D., Lai, M-C.D., and Sick, V., "The Influence of Swirl
Ratio on Turbulent Flow Structure in a Motored HSDI
Diesel Engine~ A Combined Experimental and Numerical
Study," SAE Paper 2004-01-1678, SAE Transactions,
113, No. 3, pp. 1094-1129, 2004.

17. Dec, J.E. and Zion, P.L.K. The Effects of Ignition


Timing and Diluent Addition on Late-Combustion Soot
Burnout in a DI Diesel Engine based on Simultaneous 2D Imaging of OH and Soot SAE Paper 2000-01-0238,
2000.
18. Dec, J.E. and Espey, C. Chemiluminescence Imaging of
Autoignition in a DI Diesel Engine, SAE Paper 982685,
SAE Transactions, 107, 1998.
19. Picket, L. M., Siebers, D.L., and Idicheria, C.A.
Relationship between Ignition Processes and the Lift-Off
Length of Diesel Fuel Jets, SAE Paper 2005-01-3843,
2005.

Kong, S-C. and Reitz, R. D., "Use of Detailed Chemical


Kinetics to Study HCCI Engine Combustion with
Consideration of Turbulent Mixing Effects. " J. Engng
Gas Turbines and Power, 124(3), 702-707, 2002

20. Singh, S, Reitz, R.D., and Musculus, M.P.B.


Comparison of the Characteristic Time (CTC),
Representative Interactive Flamelet (RIF), and Direct
Integration with Detailed Chemistry Models against
Optical Diagnostic Data for Multi-Mode Combustion in a
Heavy-Duty DI Diesel Engine, SAE Paper 2006-010055, 2006.

Kong, S. C., Kim, H., Reitz, R. D., and Kim, Y.,


"Comparisons of Combustion Simulations using a
Representative Interactive Flamelet Model and Direct
Integration of CFD with Detailed Chemistry," ASME
ICES2005, 2005.

21. Singh, S, Reitz, R.D., Musculus, M.P.B., Lachaux, T.,


Validation of Three Different Engine Combustion
Models Against Detailed in-Cylinder Optical Diagnostics
Data for a Heavy-Duty DI Diesel Engine, Submitted to
IJER, 2006.

10. Patterson, M. A. and Reitz, R. D., "Modeling the Effects


of Fuel Spray Characteristics on Diesel Engine
Combustion and Emissions," SAE Paper 980131, SAE
Transactions, 107, No. 3, pp. 27-43, 1998.
6

You might also like