You are on page 1of 11

2006-01-0241

Adjustment and Verification of Model Parameters for


Diesel Injection CFD Simulation
Prof. Dr. Winfried Waidmann
Fachhochschule Aalen, Aalen, Germany

Dr. Andreas Boemer


DEUTZ AG, Kln, Germany

Dr. Markus Braun


Fluent Deutschland GmbH, Darmstadt, Germany
Copyright 2006 SAE International

ABSTRACT
In co-operation between a diesel engine manufacturer
(DEUTZ), a CFD software supplier (FLUENT) and two
universities (FH Aalen and RWTH Aachen), the practicability of industrial CFD simulation for the Diesel injection process has been investigated. Since these simulations still rely on several empirical parameters, the main
aim of this investigation was to obtain an adjusted set of
model parameters in order to achieve realistic results
with the current FLUENT version.
To verify the simulations, droplet diameters and velocity
distributions of the spray have been measured by Phase
Doppler Anemometry (PDA) in a model chamber.
A Lagrangian Discrete Phase Model has been used for
spray simulation. Different primary break-up models and
two secondary spray break-up mechanisms, the KelvinHelmholtz theory and Rayleigh-Taylor instabilities have
been investigated. Furthermore, droplet evaporation,
droplet collision and the influence of droplet shape on
drag have been accounted for.

INTRODUCTION
The patented high pressure Deutz-Common-Rail Diesel
injection system, equipped with BOSCH injectors, results
in spray atomization with very small mean droplet diameters. This guarantees effective evaporation of the fuel
and enhanced mixing of the reactants. Multi-dimensional
simulation of these processes is still an issue. Although
CFD claims to rely on physical principles rather than empirical formulas, and although there exist sophisticated
theories of Diesel sprays, the spray models currently
available in literature will not work without adjustment of
constants and fitting parameters based on detailed
measurements. However, since simulation of combustion and pollutant chemistry is not possible without cor-

rectly predicting the spray, DEUTZ launched a project to


perform such an adjustment for one of their actual Diesel
engines using its CFD code FLUENT.
The Lagrangian Discrete Phase Model of FLUENT 6.2
has been applied for this purpose. Various sub-models,
accounting for the effects of turbulent dispersion, coalescence, collision, evaporation and droplet break-up were
tested and compared with experimental results.
The experiments were conducted with an original injector
mounted at the optically accessible pressure chamber of
the Lehrstuhl fr Wrme- und Stoffbertragung der
RWTH Aachen, Germany (WSA, Aachen University) [1,
2]. To avoid ignition during measuring, the temperature
and pressure inside the model chamber were kept below
self ignition conditions (710 K, 50 bars). Phase Doppler
Anemometry (PDA) was applied to obtain locally resolved droplet diameter and droplet velocity distribution.
Furthermore, flashlight shadowgraphs delivered visualizations of the time dependant development of the spray.
In this paper, first a brief review of the applied numerical
models, model parameters and general conditions is
presented. Next, the experimental setup is briefly described. In the following main part, the results calculated
with the proposed set of model parameters are presented and compared with the experimental data.

NUMERICAL MODEL
The numerical calculations were carried out with the
FLUENT 6.2 CFD code. The chosen multiphase model is
the Discrete Phase Model, applying the stochastic tracking of individual droplets (Lagrangian approach). All
droplets were injected with the same initial diameter
('blob'-method). In general, this diameter is identical to
the nozzle orifice diameter or somewhat decreased due
to cavitational or turbulent effects. In the present case,

consideration of these effects has been found not to be


necessary. The liquid phase is assumed to be an artificial single component Diesel surrogate with a H/C ratio
3
of 29/16, density 810 kg/m , thermal conductivity 0,12
W/m/K, specific heat as the following linear function of
temperature

c p [ J / kg / K ] = 835.04 + 3,73 * T [ K ]
droplet surface tension as the following polynomial function of temperature

[ N / m] = 0.0493868 8.26605E 5 * T [ K ]
1.25984E 7 * T [ K ]2 2.06675E 10 * T [ K ]3
saturation vapor pressure piecewise linear between 1329
Pa at 352 K and 50500 Pa at 600 K, viscosity piecewise
linear between 0.0047573 Pa s at 273 K and 0,0005089
Pa s at 423 K, vaporization temperature 352 K, boiling
point 453 K. These data were, if possible, derived from
measurements of real Diesel or otherwise taken from
literature. For all simulations, the standard k, turbulence model has been applied. A dynamic drag model
has been chosen, which accounts for the effects of droplet distortion by linearly varying the drag between that of
a sphere and a disk, according to the actual aerodynamic forces [3]. Second order upwind discretization has
been applied for all conservation equations. Further refinement of the computational grid and lowering the time
step size did not change the results considerably.
Table 1 gives an overview of the finally proposed physical models, model parameters and general conditions. A
brief description of the injection model and the break-up
models follows. Readers can find more detailed information in [7] or www.fluent.com.
PRIMARY SPRAY BREAK-UP
The Solid Cone Model is the only FLUENT atomizer
model that was found to lead to satisfactory results in
these simulations. All other primary break-up models or
injection types, including the should-be appropriate Plainorifice Atomizer Model, delivered unrealistic penetration
lengths and spray shapes. For the Solid Cone Model, the
spray angle has to be specified as an input value. In this
calculation the half cone angle was taken from the spray
shadowgraphs. Sensitivity studies showed that the cone
angle has a significant influence on the spray pattern. It
has therefore to be adjusted carefully in case of varying
boundary conditions. The correlation from [8],

tan = a a
f
2

0.19

f
0.0043
a

0. 5

with the spray angle , an adjustable constant a, air

density a , and fuel density

can be used for this

task. E. g., increasing the pressure from 50 bar (pres-

sure chamber condition) to 140 bar (real engine condition) will enhance the spray half angle from 10 to 12.4
degree according to this equation.
Models

Parameters

Comments

Solid cone
injection

10 degree
cone half angle
B0 = 0.61,
B1 = 18,
C3 = 2.5,
c = 30

Primary break-up,
value metered from
the shadowgraphs
Secondary break-up

Combined Kelvin-Helmholtz
(KH) and
Rayleigh-Taylor
(RT) [4], [5]
Droplet collision

Default [6]

Necessary in combination with the secondary break-up


model
Initial droplet
0.167 mm
Identical to nozzle
diameter
diameter
Fuel injection
330 K
50 K below measured
temperature
nozzle temperature
Aerodynamic
Dynamic drag Includes droplet dedrag [3]
coefficient
forming due to aerodynamic forces
Injection velocity Variable,
Calculated from
max. 430 m/s measured time dependant mass flux
(Figure 2)
Turbulent droplet Default
Turbulent tracking of
dispersion
the droplets
Number of in500 parcels
Distributes the disjected particle
per time step crete phase source
streams
terms onto the flow
Time stepping
Corresponds to 0.5
50 s
degree of crank angle
Turbulence
Standard k, - Turbulence model not
varied
model
Air density
Ideal gas
Real gas effects neglected
Table 1: proposed model parameters
SECONDARY SPRAY BREAK-UP
The secondary break-up model is the backbone of spray
simulation. The models usually applied, calculate liquid
break-up by using some form of stability analysis [4].
Break-up occurs if droplets exceed a certain stability criterion, and the characteristics of the resulting droplets
depend on the wavelength of the instability that caused
the break-up. In the present study, Kelvin-Helmholtz (KH)
and Rayleigh-Taylor (RT) models were tested. The KH
model (also called Wave-Model) considers the stripping
process of droplets due to the growth of KelvinHelmholtz instabilities on the droplet surface, resulting
from the relative velocity between the gas and the liquid
phase. The RT model accounts for sudden catastrophic
break-ups due to the deceleration of the droplets. The
resulting droplets are larger than those of the KH model.
Physical details of the different models and proposed
activation conditions are described in [5]. In accordance
with their findings, the present study is showing that a

combination of these two models, the so-called KH-RT


model, can match the measured droplet size distribution
best. In the vicinity of the nozzle, up to a certain path
length (the so-called break-up length), it is expected that
droplets undergo only KH break-up, whereas further
downstream both mechanisms seem to be present. The
break-up length can be calculated according to [5] by

Lb = c

f
d
a 0

with an adjustable constant c, the orifice diameter


and the densities of air

a and

d0 ,

fuel f . With the pro-

posed value of c = 30, the break-up length is about 30


mm.
The combined KH-RT model has been found to be an
adequate tool to simulate the secondary break-up of the
spray satisfactorily. The parameters applied in this investigation are similar to values reported in literature, see
Table 2. B0 is a multiplier for the wavelength of the fastest growing wave on the droplet surface, B1 is a parameter accounting for nozzle effects, and C3 rules the
weighting of the RT model. With the proposed value C3
= 2.5, the influence of the RT model is only weak, while
the KH model is dominating.
B1

C3

Source

18

2.5

30

This investigation

10

5.33

[9]

40

0.1

17

[10]

10

2.5

[11]

10

1.0

[10]

10

2.5

[10]

12

1,0

30

[12]

Table 2: KH-RT model parameters (B0 = 0.61)


The KH model alone leads to quite reasonable results,
but the combination with the RT model improves them
furthermore.

rithm, e. g. [13], but such models are not yet available in


FLUENT.
EVAPORATION
Although real Diesel fuel consists of a blend of hundreds
of components, it is usually treated as a single component. In the present case, the evaporating species is assumed to be C7H16 only. This is in fact an oversimplification. Multicomponent evaporation models have
been proposed in literature, e. g. [14], but are not yet
available in FLUENT as a standard model.
Temperature gradients inside the droplet have been neglected. Vaporization has been assumed to take place
as soon as the droplet temperature exceeds the boiling
temperature. Heat- and mass transfer were calculated in
the usual way from a Nusselt correlation [15] with the
actual droplet diameter. A constant binary diffusion coef2
ficient of D = 3.79 E-6 m /s and a latent heat of hv = 180
kJ/kg were applied. This is justified since the ambient
conditions in the model chamber are constant. Radiation,
supercritical effects and real-gas effects were neglected.

EXPERIMENTAL SETUP
The experimental data for the adjustment of the model
parameters were measured in the model chamber of
WSA, Aachen University [1, 2]. The main conditions are
listed in Table 3. A sketch of the chamber is shown in
Figure 1.
Common Rail Injec- Bosch B445120876
tor
Number of injection 7, only one jet is observed
holes
Injection pressure
1600 bar
Orifice diameter
0.167 mm
3
Injected Diesel vol- 154 mm
ume
Pressure and tem- Below self ignition (50 bar, 710 K)
perature
Purge flow
0.05 m/s, air
Fuel
EN 590 summer Diesel
Table 3: experimental conditions
The seven-hole-injector was mounted in such a way that
one spray was directed vertically. Only this spray was
considered, while the other ones have been deflected. A
moderate airstream from the top to the bottom of the
chamber cleaned the chamber and the observation windows between the successive test runs.

COLLISION AND COAGULATION


The commonly used collision model of [6] has been applied in this investigation. It uses a stochastic estimation
of droplet collisions and predicts either coalescence or
bouncing. The probability of each outcome is calculated
from the collisional Weber number and a fit to experimental observations. See [7] for further details. There
are newer models proposed in literature which use adaptive superimposed meshes and the NTC collision algo-

A Phase Doppler Anemometer (PDA), see [16], delivered


the droplet diameter and droplet velocity distributions.
The measurement points have been taken at three axial
distances from the injector at 40, 45, and 50 mm. The
radial measurement points from the spray center to the
spray boundary have a spacing of 1 1.5 mm. This region is suitable for the measurements, since it provides a
fully developed spray, sufficiently optical transparency

and high enough droplet numbers for a reliable data acquisition rate.
air inlet
(T = 710 K,
p = 5 MPa)

quartz glass
window

pressure
chamber
common-rail
injector

incident
beams
liner

100 mm

outlet

Figure 1: Model chamber of WSA, Aachen University


Furthermore, nanolight shadowgraphs were taken with a
high-speed camera every 50 s to document the spray
development. The light flashes (light duration 20 ns) are
absorbed and scattered depending on the local material
properties of the illuminated fluid. Thus regions containing droplets, vapor and air can be distinguished by the
grey-scale of the shadow images. More detailed information on the shadowgraph technique can be found in [17],
e. g.
The knowledge of the flow characteristics of the injected
fuel at the nozzle exit is a key issue for a successful
simulation of the spray. Figure 2 shows the measured
needle lift and the corresponding mass flux of the fuel.
The mass flux course results from the injector characteristics and the rail pressure. In the model chamber, these
conditions were kept identical to those in the real DEUTZ
engine. The time delay between needle lift and start of
injection was about 350 s. After an injection time of
2250 s, the totally injected fuel mass of one hole was
18.1 mg.

RESULTS AND DISCUSSION


Figure 3 shows a comparison of the calculated spray
pattern with shadowgraphs at t = 1000 s injection time
(1350 s after injector triggering). The three depicted
shadowgraphs are a result of short-time recordings with
an exposure time of a few nanoseconds. They have
been taken at different test runs at the same instant of
injection, but show a strongly irregular spray pattern, especially at the tip region. The overall fuel penetration
length is nearly identical in all three shadowgraphs, but
the tip region sometimes consists of droplets (dark black
region), sometimes of vapor (grey regions). The measured conical spray boundary also shows fluctuations with
small outbursts of droplets or vapor. Compared to the
experiments, there are much less irregular coves visible
at the simulated jet. This is due to the applied ReynoldsAveraged Navier-Stokes (RANS) equation set. Nevertheless, penetration length and overall shape of the spray
are well reproduced. Note that the cone angle for the
primary break-up model has been taken as a timeaverage from the experiments.
The calculated droplet distribution shows a dense core
region near the orifice. Droplets at the jet boundaries are
more exposed to the surrounding air and experience
stronger deceleration than the droplets located in the
central jet region. Hence, about 20 mm away from the
nozzle, radial spray expansion starts. Vapor mass fraction is almost uniformly distributed in the spray region,
only close to the tip a maximum vapor mass fraction of
about 0.3 occurs. The overall shape of the vapor distribution is nearly identical to the droplet distribution. In the
real engine vapor will be transported by the swirl in the
combustion chamber. At the tip the vapor elongates a
few millimeter further into the surrounding air medium.

Figure 3: Calculated droplet distribution,


shadowgraphs (black = liquid, dark grey = vapor, light
grey = air), vapor mass fraction distribution
(blue = 0, red = 0.3)

Figure 2: Injector needle lift and mass flux course

A comparison of the complete injection process is given


in Figure 4. The shown shadowgraphs are an average of
30 single shots taken from different test runs at analogous instants. The injection starts at t = 0 s (350 s after
the injector needle lift) and ends at t = 2250 s, when the
droplet cloud lifts off from the injector orifice. At t = 2850
s, the droplet cloud is almost evaporated. The calcu-

lated time dependant droplet penetration length and the


overall shape of the spray are in good agreement with
the experiment. Sharp spray tips, clear vapor/air or droplet/vapor interfaces, like visible in Figures 3 and 7, can-

not be observed in these images because of the averaging.

Figure 4: Comparison of averaged shadowgraphs (black = liquid, dark grey = vapor, light grey = air) with the calculated
droplet distribution (blue = 310 K, red = 453 K droplet temperature)

Figure 5 shows the development of the calculated total


droplet and vapor mass corresponding versus time. The
solid upper line, the sum of droplets and vapor, is identical to the integral of the mass flux course shown in Figure 2.

After about 1100 s, the experimental data show a


strong increase of scattering and a nearly constant or
somewhat decreasing mean penetration length, whereas
the calculated spray still continues to elongate. However,
only very few droplets have been observed to fly so far in
the simulation. In the experiment, such isolated droplets
could also be present, but were probably not detected. A
reason for the increased scattering of measured data
could be the presence of huge amounts of fluctuating
vapor at the jet tip. The analysis of the experiments was
therefore performed until t = 1000 s only, and for adjustment of the model constants most attention was
turned to droplet diameter and velocity distributions.
DROPLET DISTRIBUTION

Figure 5: Calculated droplet and vapor mass during the


injection
In the beginning, droplet and the total mass are nearly
the same while the amount of vapor mass is negligible.
Evaporation begins to start at about t = 400 s. Subsequently, the vapor mass fraction increases nearly linearly
with time until the end of injection. At t = 2500 s the fuel
is completely evaporated.

In order to reach a sufficient data acquisition rate, PDAmeasurements can be performed in relatively dilute regions only. Suitable points were found at y = 40, 45 and
50 mm distance from the nozzle orifice where the droplet
concentration is relatively low. The radial coordinates
cover a range of 7 mm with a spacing of 1 1.5 mm.
Figure 7 shows a sketch of the axial and radial measurement positions.

The slope of the vapor mass fraction is almost parallel to


the totally injected mass. This means the mass of droplets in the chamber remains nearly constant during this
time period, and the fuel evaporation rate is proportional
to the injected mass flow rate. This tendency can also be
observed in the experiment, where it was found that during this time period the droplet distribution and the droplet diameter histograms are more or less time independent. Fortunately, this makes droplet sampling easier.
Sampling over a certain time window is necessary to collect enough data for the construction of the droplet diameter histograms.

Figure 7: Left: shadowgraph, right: simulated droplets.


Positions for data acquisition (y = 40, 45, 50 mm;
r = 18 mm).

Figure 6: Spray penetration


Figure 6 shows the time dependant penetration length of
the liquid phase. For t < 1000 s the calculated penetration is in very good agreement with the data. The overall
agreement is better than 5 percent.

Since the injection process is strongly time-dependant,


comparison of the experimental data with the simulation
results requires summarising and averaging of the data
over a certain time. Fortunately, there is a period when
the mean droplet diameters are nearly constant. This can
be observed in Figure 8, where the experimental mean
droplet diameter is plotted versus time. Between 1.3 and
2.5 ms, the diameters at the different radial positions do
not vary significantly. Hence, this period has been chosen for obtaining the data. In the simulation, the droplets

were sampled in a ring of 1 mm thickness. Smaller


thicknesses led to a decrease of statistical accuracy.

diameter, which is an important parameter for the


evaporation model.

The experimental data indicate that the mean droplet


diameter decreases with a factor of about two from the
core region r = 0 mm to the outer spray boundary at r =
7.5 mm. This tendency has also been observed in the
simulation, and can be explained with droplet evaporation.

The presence of unrealistic large droplets can also be


observed at y = 45 and 50 mm in the simulations. An
explanation of this artefact was given in [10] on the basis
of the applied KH break-up model. According to this
model, droplets do not completely break up into new
droplets in a time step, but only shrink. The stripped
material forms a growing number of small droplets, but
the original droplets do not disappear completely. The
RT model could theoretically improve this behaviour
because it assumes sudden break-up of the complete
droplet, but increasing its influence by decreasing the
parameter C3 has been found to result in an even worse
agreement with the experiments.

DROPLET DIAMETER DISTRIBUTION HISTOGRAMS


Figure 9 shows a comparison of the calculated and
measured droplet diameter distribution for several radial
positions at y = 40 mm distance from the injector. At the
spray center (r = 0 - 1 mm) the maximum of the
calculated droplet histogram is shifted to larger
diameters in comparison to the measured ones. For
greater radial distances, the calculated maximum is
shifted to smaller diameters. At r = 2 - 3 mm, the maxima
of the simulated and measured spectra agree very well.

Figure 8: Measured mean droplet diameter versus time


at y = 50 mm
A growth of the KH parameter B1 for example from 18 to
20 would result in larger droplets and hence a better
agreement of the droplet maximawould be achieved.
Nevertheless, the value B1 = 18 has finally been chosen
because of a better agreement of the velocity
distribution.
For radial positions above r = 6 mm, the calculated
droplet histograms do not show a clearly defined
maximum. The variance of the calculated diameters is
considerable in this region. This is caused by the relative
low number of droplets at the border area of the spray.
All calculated droplet distributions show a sort of bimodal
character. There exists a distribution of smaller droplets
which corresponds well with the measured data, but
there are also bigger ones which have not been
observed in the experiment. This effect degrades the
compliance of the mean diameters, especially the Sauter

Figure 10: Mean droplet diameters


.

Figure 9: Droplet size distribution comparison of the CFD simulation results with measured data at different positions
(values in brackets: mean droplet diameter)

Figure 11: Droplet velocity histograms (values in brackets: mean velocities)

The mean droplet diameters estimated from all


histograms are given in Figure 10. They are
overpredicted due to the described presence of some
large droplets. The tendency of decreasing droplet
diameter with increasing distance from the spray axis is
well predicted in the simulation. In the outer region of the
spray (r > 6 mm), the calculated values show a higher
scattering and the mean diameters differ slightly from
this tendency.

VELOCITY HISTOGRAMS
Figure 11 shows comparisons of the calculated and
measured velocity histograms for the positions y = 45
mm and r = 0 7 mm. The calculated histograms agree
very well with the measured ones. In the outer region of
the spray (r > 5 mm), the shape of the calculated droplet
distribution became narrower than the experimental one.
The mean droplet velocities evaluated from each

histogram are shown in Figure 12 and agree very well


within a mean deviation of about 3.5 percent. Due to the
density ratio of about 30, and the discontinuous
consistency of the liquid phase, the observed linear
velocity decrease differs from what one would expect
from a single phase free jet.

FLUENT, described in this paper, a homogeneous


droplet temperature is assumed. From Figure 13 it could
be concluded that in contrast to the situation in the model
chamber, in the real engine the droplets reach the critical
state. In engine conditions, about 58 % of the droplet
mass evaporate in the supercritical state according to
this calculation. If n-dodecane, another popular Diesel
surrogate, is used, the corresponding value would even
be about 85 %. However, in a real spray the surface
tension tends towards zero before reaching the critical
state. Aerodynamic forces will then cause droplet breakup and the droplet will evaporate in an under-critical state
[18]. For the purpose of this investigation, supercritical
evaporation effects had to be neglected.
1,2

Diameter (d/d0)

1
0,8

650
K
600
550

critical
point

0,6

500

0,4

450
Model Chamber
( 710 K, 50 bar)

0,2
0
0

0,02

0,04

0,06
Time

400

Surface Temperature

Engine
(1000 K, 150 bar)

350
0,08 ms 0,1

Figure 13: Calculated droplet diameter and temperature


(n-heptane, d0 = 8 m)
The investigations presented in [18] led to the
conclusion, that at the high pressures and temperatures
prevailing in a Diesel engine real gas effects have to be
accounted for. However, the detected influence on the
droplet diameter distribution was rather marginal.
Furthermore, with increasing pressures the consideration
of Diesel as a multicomponent fuel becomes even more
important for predicting evaporation [14]. This has been
neglected in this investigation, too.

TRANSFERABILITY TO REAL ENGINE CONDITIONS

Currently there is no alternative to presuming that the


model parameters adjusted by means of the model
chamber experiments are also suited for real engines.
However, some reservations remain.

The ambient conditions in the model chamber, 710 K


and 50 bar, differ from those in a real Diesel engine,
where about 1000 K and 150 bar can be expected at the
end of compression.

In the meantime, the proposed set of parameters has


been successfully applied to simulate combustion and
pollutant emission of a DEUTZ Diesel engine. Results
were presented in [19].

Figure 13 shows diameter and surface temperature


history of a single n-heptane droplet moving with 100 m/s
in either model chamber or engine conditions. Its initial
diameter is 8 m, which corresponds to a representative
mean diameter observed at the measurements (see
Figure 8). The calculations were carried out at WSA,
Aachen University, with a model similar to the one
described in [18]. It takes a temperature profile inside the
droplet into account, whereas in the simulations with

CONCLUSIONS

Figure 12: Mean droplet velocity distribution

Based on experiments carried out at a model chamber,


models and parameters for CFD-simulation of commonrail Diesel injection have been adjusted.
The model parameter set obtained by this means is able
for accurate calculation of the penetration length of the
spray (mean deviation 5 percent), the local droplet size

(mean deviation 15 percent) and droplet velocity (mean


deviation 3.5 percent). The experimentally observed
decreasing droplet diameter with increasing radial
distance can also be predicted correctly. The diameter
distribution was the most sensitive value during the
adjustment process, whereas the simulated velocity
distribution and the penetration length matched the
experiments with various sets of parameters. Further
improvement of the break-up and coagulation models is
therefore desirable.
Although the proposed set of parameters has been
successfully applied to simulate combustion and
pollutant emission, indications exist that supercritical,
multicomponent evaporation and real gas effects should
be taken into account additionally.

ACKNOWLEDGMENTS
The contribution of the WSA, Aachen University, especially by Dr. Michael Staudt, is gratefully acknowledged.
For further information on their measurement techniques
and theoretical work see www.wsa.rwth-aachen.de.

REFERENCES
[1] M. Staudt, A. Pawlowski; Strmungsuntersuchungen
mit optischer Messtechnik; Lehrstuhl fr Wrme- und
Stoffbertragung, Aachen University; Ergebnisbericht
Deutz AG; 2004.
[2] M. Staudt, U. Meingast, U. Renz; Experimental Investigation of the Turbulent Flow Field of Vaporizing Diesel
Sprays; Proc. of the 2002 Fall Technical Conference of
the ASME Internal Combustion Engine Division: Design,
Application, Performance and Emissions of Modern Internal Combustion Engine Systems and Components;
ICE-Vol. 39, 08.-11.09.2002, New Orleans, Louisiana,
USA; 289/297
[3] A. B. Liu, D. Mather, R. D. Reitz; Modeling the Effects
of Drop Drag and Breakup on Fuel Sprays; SAE Technical Paper 930072, 1993
[4] R. D. Reitz; Modeling Atomization Processes in High
Pressure vaporization Sprays; Atomization and Spray
Technology 3, 1987.
[5] M. A. Patterson, R. D. Reitz; Modeling the Effects of
Fuel Spray Characteristics on Diesel Engine Combustion
and Emission; SAE paper 980131, 1998.

[6] P. J. ORourke; Collective Drop Effects on Vaporizing


Liquid Sprays, Ph.D. Thesis, Princeton University, 1981.
[7] FLUENT User's manual, Version 6.2, 2005
[8] J. D. Naber, D. L. Siebers; Effects of Gas Density and
Vaporization on Penetration and Dispersion of Diesel
Sprays, SAE 960034; 1996.
[9] L. Alloca, P. Belardini, C. Bertoli, F. Corcione, F. De
Angelis; Experimental and numerical Analysis of a Diesel
Spray; SAE Paper 920576; 1992.
[10] C. Baumgarten, S. Pttker, G. P. Merker; Simulation
der Spraybildung in direkteinspritzenden Dieselmotoren,
Simulation der motorischen Verbennung; Haus der
Technik, 2004.
[11] T. Kamimoto; H. Yokota; H. Kobayashi; Effects of
High Pressure Injection on Soot Formation Processes in
a Rapid Compression Machine to Simulate Diesel
Flames; SAE Paper 871610; 1987.
[12] C. v. Knsberg Sarre, R. Tatschl; Spray Modeling/Atomization Current Status of Break-up Models;
Imech Turbulent Combustion of Gases and Liquids
Leading Edge Technologie, Two Day Seminar, 1998,
The Wawn, Lincoln, U.K.
[13] S. Hou, S. Are, D. P. Schmidt; A Generalized Adaptive Collision Mesh for Multiple Injector Orifices; ERC's
th
13 International Multidimensional Engine Modeling
User's Group Meeting; Detroit, USA; 2003.
[14] Y. Yi, G.-S. Zhu, R. D. Reitz; High-pressure Spray
Combustion Modeling using Continuous Thermodynamics for Diesel Fuels; SAE paper 2001-01-0998, 2001.
[15] W. E. Ranz, W. R. Marshall Jr.; Evaporation from
Drops, Part I.; Chem. Eng. Prog., 48(3):141-146, March
1952.
[16] H. E. Albrecht, M. Borys, N. Damasche, C. Tropea;
Laser Doppler and Phase Doppler Measurement Techniques; Springer Verlag Berlin Heidelberg New York;
ISBN 3-540-67838-7; 2003.
[17] G. S. Settles; Schlieren and Shadowgraph Techniques, Springer Verlag Berlin Heidelberg New York; ISBN
3-540-66155-7; 2001.
[18] S. Hohmann; Strahlausbreitung und Tropfenverdunstung bei der dieselmotorischen Direkteinspritzung;
PhD thesis, Aachen University; Shaker Verlag, Aachen,
Germany; ISBN 3-8265-6246-1, 1999.
[19] D. Gonsior, A. Boemer, W. Waidmann; CFD Simulation des Verbrennungsprozesses in einem DEUTZDieselmotor; Fluent CFD-Forum 2005; Bad Nauheim,
Germany.

You might also like